Next Article in Journal / Special Issue
Influence of Process Parameters on Porosity and Hot Cracking of AISI H13 Fabricated by Laser Powder Bed Fusion
Previous Article in Journal
Microstructural and Micromechanical Characterization of Feedstock for Cold Spray Additive Manufacturing and Processing
Previous Article in Special Issue
Understanding Morphology, Bulk Properties, and Binding of Silk Particles for 3D Printing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Microfluidic Synthesis of Solid Catalysts

1
School of Chemistry and Chemical Engineering, Henan University of Technology, Zhengzhou 450001, China
2
Department of Chemical Engineering, Engineering and Technology Institute Groningen, University of Groningen, 9747 AG Groningen, The Netherlands
*
Author to whom correspondence should be addressed.
Powders 2022, 1(3), 155-183; https://doi.org/10.3390/powders1030011
Submission received: 19 June 2022 / Revised: 26 July 2022 / Accepted: 1 August 2022 / Published: 4 August 2022
(This article belongs to the Special Issue Feature Papers in Powders)

Abstract

:
Heterogeneous catalysis plays a central role in the chemical and energy fields, owing to the high and tunable activities of solid catalysts that are essential to achieve the favorable reaction process efficiency, and their ease of recycle and reuse. Numerous research efforts have been focused on the synthesis of solid catalysts towards obtaining the desired structure, property and catalytic performance. The emergence and development of microfluidic reactor technology provide a new and attractive platform for the controllable synthesis of solid catalysts, primarily because of its superior mixing performance and high heat/mass transfer efficiency. In this review, the recent research progress on the synthesis of solid catalysts based on microfluidic reactor technology is summarized. The first section deals with the synthesis strategies for solid catalysts, including conventional methods in batch reactors and microfluidic alternatives (based on single- and two-phase flow processing). Then, different kinds of solid catalysts synthesized in microflow are discussed, especially with regard to the catalyst type, synthetic process, structure and property, and catalytic performance. Finally, challenges in the microreactor operation and scale-up, as well as future perspectives in terms of the synthesis of more types of catalysts, catalyst performance improvement, and the combination of catalyst synthesis process and catalytic reaction in microreactors, are provided.

Graphical Abstract

1. Introduction

Solid catalysts have attracted a tremendous amount of research interest due to their wide range of applications in science and industry [1,2]. According to statistics, more than 80–90% of industrial chemical processes use catalysts to work effectively, and the vast majority of produced chemical products and commodities involve catalytic conversions [3,4]. Moreover, the research hotspots in recent decades, such as water purification, fuel cell, solar cell, hydrogen generation and storage, biodiesel production, drug delivery and release, are inseparable from the application of catalysts [5]. Nowadays, catalysts have become the core of the entire chemistry, chemical engineering, energy and environmental industries [6].
The emergence of nanotechnology has brought new opportunities to the research and evolution of solid catalysts [7]. Micro- and nanoscale catalysts have been found to possess better catalytic properties than their macroscale counterparts [8,9]. When the size of catalytic materials decreases to micrometer or even nanometer level, its specific surface area, surface steps, wrinkles and defects increase, resulting in the increase in catalyst active sites, which can greatly improve the catalytic activity and selectivity [10,11]. Accordingly, the preparation of micro- and nanoscale catalytic materials can improve the application efficiency of catalysts to a large extent. In addition, the preparation method significantly affects the crystal structures (e.g., shape, size, composition) and texture properties (e.g., porosity, specific surface area) of catalysts, which has a great impact on the catalytic performance. Therefore, the controllable synthesis of catalysts is crucial to enhance the efficiency and sustainability of catalytic processes [12].
Chemical precipitation method is one of the most commonly used methods for the preparation of catalytic materials (e.g., non-precious metal catalysts [13], metal oxide catalysts [14,15,16,17] or supports [18]). The precipitation process in laboratory or industrial production is usually carried out in batch stirred tanks. Due to the simple operation and low technical requirements of the precipitation process in batch, micro- and nanoscale solid catalysts with various morphologies and structures (e.g., nanospheres, nanorods, nanoplates, nanorings and microflowers) have been synthesized. However, because of the inherent process characteristics of batch operation mode, the mixing rate, mass and heat transfer efficiency are usually low, often resulting in a wide particle size distribution of the obtained catalytic materials and poor batch-to-batch repeatability [19]. In addition, the inability to precisely control the reaction time of batch reactors results in the overlapping of the reaction precipitation process with the subsequent structural rearrangement process. Namely, the nucleation, growth, and agglomeration processes may occur simultaneously [20]. This renders the inhomogeneous particle size, morphology and microstructure of catalysts obtained by the precipitation method in batch, which is not conducive to the well-controlled synthesis and large-scale production of catalytic materials [21].
The emergence and development of microfluidic reactor technology as a typical chemical process intensification method offer great opportunities to overcome the most pressing drawbacks of the precipitation method in batch as well as other conventional methods for the synthesis of solid catalysts [22]. This technology uses microchannels (with lateral dimensions typically from tens to hundreds of micrometers) to precisely manipulate and process fluids in a continuous flow manner [23,24,25]. In such confined microchannels, the specific surface area is significantly increased and the diffusion/conduction distance is much reduced. This, combined with the well-regulated laminar flow pattern under operation, renders an accelerated mixing speed and improved heat/mass transfer efficiency, which is beneficial to the reaction process control and regulation of the properties of the synthesized particles [26,27]. Microfluidic reactor technology thus offers great development opportunities in the field of synthetic chemistry in both academia and industry, including the controllable synthesis of solid catalysts as well [28,29]. Up to now, various solid catalyst particles have been successfully synthesized in microflow, such as noble metal nanoparticles (Au, Ag, Pd, Pt) [30,31,32,33], bimetallic nanoparticles [34,35], supported-metal nanoparticles [36], metal oxide nanoparticles (ZnO, CeO2, TiO2) [37,38,39], zeolites [40], metal-organic frameworks (MOFs) [41,42] and so on.
In this article, we briefly review the latest research progress on the utilization of microfluidic reactor technology for the preparation of solid catalysts. The review is divided into three sections. First, the synthesis principles of solid catalysts, including the underlying synthesis process, the commonly used conventional methods and microfluidic alternatives (based on single- and two-phase flow processing), are summarized. The following section addresses various types of solid catalysts synthesized by microfluidic methods, and the comparison of catalyst structure and performance prepared by different methods. Finally, we discuss the challenges in the microreactor operation for solid handling and scale-up, the integration with process analytic tools, as well as perspectives in the improvement of catalyst performance, the synthesis of new catalysts, the combination of synthesis process and catalytic reaction in microreactors. This review mainly focuses on the different microfluidic synthesis methods, and performance comparison between the obtained catalysts. For a more extensive overview on the in situ characterization or promising reaction application of solid catalysts using microfluidic reactors, and other application examples of solid catalysts synthesized in microflow, we refer the reader to other review articles [11,43,44,45,46].

2. Solid Catalyst Synthesis in Batch and Microflow

2.1. Synthesis Principle

The synthesis of (nanoscale) solid catalysts is largely based on wet chemical methods both in batch reactors and in microfluidic synthesis fields, which includes three stages: firstly, the reactant molecules are (uniformly) mixed (i.e., the micromixing process); secondly, a chemical reaction occurs to form a supersaturated solution; thirdly, the precipitate is formed in the crystallization process including nucleation, growth and agglomeration (Figure 1) [47]. The precipitate becomes the solid catalyst typically after filtering, washing, drying, and calcination/reduction process. When the concentration of the target product reaches supersaturation in the solution, it will be deposited by homogeneous or heterogeneous nucleation. During the precipitation process, the nucleation, growth and agglomeration of crystals have a great influence on the product particle size and morphology [48]. In order to generate non-agglomerated particles with a narrow particle size distribution, it is required that all nuclei are formed nearly simultaneously, and no secondary nucleation or particle agglomeration occurs during the subsequent crystal growth process. The particle size and its distribution and morphology of the target product depend on the relative rates of nucleation, crystal growth and agglomeration kinetics [49], which are therefore influenced by reaction conditions (temperature, concentration, solvent composition, pH, etc.) [50,51]. The morphology, size and structure have a significant impact on the physical and chemical properties of catalyst particles, which in turn affect their catalytic performance [15]. Therefore, it is of great significance to study the preparation method of catalysts to improve their catalytic activity and selectivity.

2.2. Batch Processes

The commonly used methods for solid catalyst synthesis in laboratory and industry mainly include chemical precipitation method, hydrothermal method, solvothermal method, sol-gel method and thermal decomposition method, etc. Generally, these methods are operated in batch mode, and Table 1 summarizes some solid catalysts obtained by these batch processes.
The precipitation method is the most frequently used one, which uses the chemical reaction between a salt solution and a precipitant solution to form a water-insoluble salt (Figure 2a) [52]. For example, ZnO nanoparticles were synthesized by the low temperature chemical co-precipitation between Zn(ac)2·2H2O and NaOH aqueous solution, which can be used as an effective catalyst for the degradation of Congo red dye [53]. Carbon nanotube (CNT)-Cu2O nanocomposites were successfully prepared via a facile chemical precipitation method by adding ascorbic acid solid and NaOH solution to the mixed solution of carboxylated carbon nanotubes and CuCl. The nanocomposites exhibited excellent catalytic performance in the reduction of p-nitrophenol [54]. As a catalyst for NO oxidation, MnOx-CeO2 could be obtained by the co-precipitation of the mixture of Ce(NO3)3·6H2O and Mn(CH3COO)2·4H2O with (NH4)2CO3 solution [55], as summarized in Table 1.
In the hydrothermal method (Figure 2b) [56], a precipitant or chelating agent is added to the salt aqueous solution to prepare the precursor. Then, the mother liquor containing the precursor is poured into a closed vessel. Under the reaction conditions of high temperature and high pressure in a closed system, the insoluble substances are dissolved and recrystallized by using air in the kettle or the added specific oxidant. Many inorganic functional materials, metal oxide and composite oxide nanomaterials have been successfully prepared by this method (Table 1). For instance, Ni-Ce-ZrOδ catalysts were obtained via a one-pot hydrothermal method by mixing the Ni/Ce/Zr metal salt mixture solution and urea solution and keeping at 105 °C for 50 h. The catalyst exhibited a good catalytic activity and CH4 selectivity for CO2 methanation [57]. Mn-Ce oxides, which can be used as catalysts for the oxidation of toluene, were successfully synthesized via the combination of redox-precipitation and hydrothermal approach (i.e., by mixing KMnO4 and Ce(NO3)3·6H2O mixture solution with HCl solution and keeping at 100 °C for 12 h) [58]. FeS2 microparticles were synthesized by mixing FeSO4·7H2O, Na2S2O3·5H2O and elemental sulfur and keeping at 200 °C for 12–36 h via a hydrothermal method. The microparticles can be used as a heterogeneous electro-Fenton catalyst to decompose H2O2 and thus degrade diclofenac sodium [59].
The solvothermal method utilizes the principle similar to the hydrothermal method to prepare materials that cannot grow in aqueous solution or are susceptible to oxidation, hydrolysis or water sensitivity (Figure 2b). Typical solvents used are alcohols, benzene, acids, and amines. As exemplified in Table 1, Co3O4/Co(OH)2 hybrids were synthesized via a facile solvothermal method by heating the mixture of Co(NO3)2·6H2O, sodium oleate, water, ethanol, n-hexane, and KOH at 200 °C for 12 h. The hybrids exhibited superior activity for photocatalytic water oxidation [60]. Near spherical-shaped TiO2 nanoparticles with uniform size distribution were synthesized by the solvothermal method using titanium tetraisopropoxide as precursor, ethanol as solvent, and H2SO4 as hydrolyzing agent. The obtained TiO2 nanoparticles can be used as a photocatalyst for the degradation of methyl orange and methylene blue dyes [61]. The monodisperse PtCu dodecahedral nanoframes, as a promising catalyst in hydrogen evolution reaction, were obtained by a simple solvothermal method, where Pt(acac)2 and CuCl2 were used as precursors, cetyltrimethylammonium chloride as structure-director and dispersing agent, oleylamine as reductant, allantoin as co-reductant and co-structure directing agent [62].
The sol-gel method uses metal alkoxides or inorganic compounds (e.g., nitrates, chlorides, acetates) as raw materials to carry out a series of hydrolysis and condensation reactions to form a stable sol (Figure 2c) [63]. The sol undergoes aging, colloidal particle aggregation, or polymerization to form a gel. Then, the gel is dried, sintered and solidified to obtain metal oxides or metals. As summarized in Table 1, when the solutions of Cu(NO3)2·3H2O and CNTs (both in ethanol) were separately added to the mixture of tetrabutyl-orthotitanate, benzyl alcohol, ethanol and water, Cu doped TiO2-10 wt.% CNT nanocomposite powder was successfully prepared by the sol-gel method. The nanocomposite can be used as a photocatalyst for the degradation of methylene blue [64]. The mixture of carbon spheres, MoCl5, and diphenyl diselenide was ultrasonically dissolved in excessive ethanol to synthesize C@MoSe2 core-shell nanocomposites by a simple sol-gel approach, and the nanocomposite can be used as an electrocatalyst for the hydrogen evolution reaction [65]. The mesoporous Ni-M/Al2O3 (M: Fe, Co, Zr, La and Cu) nanoparticles were prepared by a one-pot sol-gel method using Ni(NO3)2·6H2O, Al(NO3)3·9H2O and M(NO3)x·yH2O as metal precursors, ethanol and propylene oxide as the solvent and gelation agent, respectively. The nanoparticles were shown as a promising catalyst in carbon dioxide methanation reaction [66].
In addition, the thermal decomposition method features the decomposition of the precursors (e.g., metal hydroxides, metal salts; usually obtained in a previous step from precipitation or crystallization in solution in batch) at a certain temperature to obtain metals or metal oxides (e.g., in the calcination stage) as catalysts. The obtained product generally maintains the morphology of the precursors. Co3O4 hexagonal nanoplates were successfully fabricated through the simple thermal decomposition of β-Co(OH)2 nanoplates (previously obtained in the precipitation process by adding CoCl2 and NaOH aqueous solutions into water), and exhibited a high catalytic activity in the reduction of p-nitrophenol [67]. Mesoporous CeO2 nanocrystals could be obtained by thermal decomposition of Ce-MOF (previously prepared via one-step hydrothermal synthesis), which had a superior catalyst activity for benzene combustion [68]. The magnetic porous Mn/Fe3O4 cubes were also synthesized by thermal decomposition of Fe-Mn-oxalate precipitant (previously obtained by adding FeSO4·7H2O and MnSO4·H2O to the oxalate solution), and this binary oxide showed a good catalytic activity and stability during the degradation of bisphenol A (Table 1) [69].
Figure 2. Typical steps involved in the commonly used solid catalyst synthesis methods in batch. (a) precipitation method; (b) hydrothermal/solvothermal method; (c) sol-gel method. (ac) were reproduced/adapted with permission from references [52,56,63], respectively.
Figure 2. Typical steps involved in the commonly used solid catalyst synthesis methods in batch. (a) precipitation method; (b) hydrothermal/solvothermal method; (c) sol-gel method. (ac) were reproduced/adapted with permission from references [52,56,63], respectively.
Powders 01 00011 g002
Table 1. The synthesis of solid catalysts by batch processes.
Table 1. The synthesis of solid catalysts by batch processes.
Catalytic MaterialSynthesis MethodReagentsSynthesis ConditionsSize/MorphologyTested Catalytic ReactionRef.
Ni/CeO2PrecipitationNi(NO3)2·6H2O, (NH4)2Ce(NO3)6, Na2CO3, NaOHpH = 10CeO2 (8–12 nm)Methane oxidative reforming[70]
ZnO PrecipitationZn(ac2)·2H2O, NaOH60 °C, 2 hNanoparticles
(≈18 nm)
Congo red dye degradation[53]
MnOx-CeO2PrecipitationCe(NO3)3·6H2O, Mn(CH3COO)2·4H2O, (NH4)2CO33 h aging at room temperature; 3 h calcination at 350/450 °CNanoparticles
(8.8–12.1 nm)
NO oxidation[55]
CNT-Cu2OPrecipitationCarboxylated carbon nanotubes, CuCl, ascorbic acid, NaOH 1 h stirring at 30 °CNanoparticlesp-Nitrophenol reduction[54]
Mn-CeHydrothermalKMnO4, Ce(NO3)3·6H2O, HCl, Ti-foil100 °C for 12 h;
4 h calcination at 350 °C
Microspheres
(2–5 μm)
Toluene oxidation[58]
Au-TiO2HydrothermalSodium citrate, HAuCl4·4H2O, Ti(OC4H9)43 h aging;
180 °C for 7 h
Microspheres (≈1.2 µm)Formaldehyde decomposition[71]
Ni-Ce-ZrOδHydrothermalUrea, metal salt solutions50 h aging at 105 °C;
4 h calcination at 500 °C
CO2 methanation[57]
FeS2HydrothermalFeSO4·7H2O, Na2S2O3·5H2O, sulfur200 °C for 24 hFramboids
(2.61–3.9 μm)
Diclofenac sodium degradation [59]
AgSolvothermalAgNO3, N,N-dimethylformamide, poly(vinylpyrrolidone)140 °C for 8 h Truncated triangular nanoplates
(200 ± 20 nm)
Styrene oxidation [72]
Au-SnO2SolvothermalSnCl4·5H2O, gold chloride, hydrazine30 min stirring; 100 °C for 18 h NanoparticlesRhodamine B degradation[73]
TiO2SolvothermalTitanium tetraisopropoxide, ethanol, H2SO490 °C for 12 h;
calcination at 400/450/500 °C
Nanoparticles (near spherical; ≈5 nm) Methyl orange degradation[61]
Co3O4/Co(OH)2SolvothermalCo(NO3)2·6H2O, KOH, sodium oleate, ethanol, n-hexane200 °C for 12 h Nanocubes (≈14 nm) Photocatalytic water oxidation[60]
Al-doped ZnOSol-gelZn(CH3COO)2·2H2O, AlCl3·6H2O, ethanol, diethanol amine2 h stirring at room temperature;
1 h calcination at 400 °C
NanoparticlesMethylene blue dye degradation [74]
Ni-M/Al2O3 (M: Fe, Co, Zr, La and Cu)Sol-gelNi(NO3)2·6H2O, Al(NO3)3·9H2O, M(NO3)X·yH2O, ethanol, propylene oxide45 min aging at room temperature;
48 h drying at 85 °C; 3 h calcination at 700 °C
NanocrystallineCarbon dioxide methanation[66]
Cu doped TiO2-CNTSol-gelTetrabutyl-orthotitanate, ethanol, benzyl alcohol, multi-walled carbon nanotubes, Cu(NO3)2·3H2O1 h stirring at 0 °C; ultrasonication for 10 min; 1 h calcination at 400 °CNanocubes (≈25 nm)Methylene blue degradation[64]
C@MoSe2Sol-gelMoCl5, diphenyl diselenide, ethanol0.5 h ultrasonication; 2 h heating at 700 °C under Ar/H2Core-shell nanoparticles (≈200 nm)Hydrogen evolution reaction[65]
Mn/Fe3O4Thermal decompositionFeSO4·7H2O, oxalic acid, MnSO4·H2O1 h calcination at 400 °C45.9 nmBisphenol A degradation[69]
Fe3O4-CuThermal decompositionFe(acetylacetonate)3, Cu(acetylacetonate)2, oleylamine, benzyl ether90 min reflux at 110 °C under stirring; 60 min heating at 225 °CNanocrystals4-Nitrophenol degradation[75]
Co3O4 Thermal decompositionCoCl2·2H2O, NaOH1 h aging at 50 °C under N2; 2 h calcination at 300 °CHexagonal nanoplates (≈100 nm)p-Nitrophenol reduction[67]
CeO2Thermal decompositionTrimesic acid, 4,4′-bipyridine, dimethyl formamide, Ce(NO3)3·6H2O24 h heating at 120 °C, 2 h calcination at 400/500 °CNanocrystalsBenzene combustion[68]
Although the above wet chemical methods performed in batch stirred tanks have the advantages of simple operation and low technical requirements, their inherent limitations tend to hinder their wide applications in both academia and industry. The conventional batch synthesis processes usually require bulky instruments or expensive autoclaves and complex stepwise operations, associated with non-optimized or low efficiency in the consumption of reagent, power and labor, and require long synthesis time. Moreover, because of its low heat transfer efficiency, for strongly exothermic reactions, the heat may not be dissipated in time especially when large amount of reagents are used in batch. This may cause an instantaneous increase in temperature and pressure, resulting in potential safety concerns. In addition, the low mixing rate and heat/mass transfer efficiency of batch reactors render temperature and concentration gradients during the synthesis process. Thus, the (nano)particle formation process cannot be precisely controlled, resulting in a large particle size, wide particle size distribution, inhomogeneous morphology and poor batch-to-batch repeatability.

2.3. Microfluidic Approaches

2.3.1. Advantages of Microfluidic Reactor Technology

For all the above-mentioned methods of solid catalyst synthesis, a good mixing between reagents in the solution and a fine control over other reaction parameters (e.g., concentration, temperature and reaction time) are important in determining the obtained catalyst structure, property and performance. Thus, there is an incentive to switch from batch to continuous flow processing in microreactors for process improvement. Numerous researches have reported the advantages of microfluidic reactor technology over conventional batch operation for the synthesis of nanoparticles using wet chemical methods [76,77]. When the precipitation process is conducted in a miniaturized microreactor, it is a facile, versatile and reproducible method due to the continuous operation mode, superior mixing efficiency and well-controlled residence time [78]. The precipitation process in microreactors can effectively differentiate the nucleation and growth processes, as well as mitigate the problem of nanoparticle agglomeration often encountered in a stirred tank, so that nanoparticles obtained in microreactors usually have a smaller particle size and more uniform particle size distribution than in batch [14]. It takes typically seconds to tens of minutes to synthesize solid catalyst nanoparticles by hydrothermal or solvothermal method in the microreactor, while it usually takes hours to days in a batch reactor [40,79]. Similarly, the synthesis of catalyst nanoparticles by the sol-gel method in a microreactor usually requires a short reaction time (e.g., 1 min), and the obtained nanoparticles have a small particle size, narrow and controllable particle size distribution [12,80].
From the discussion above, we can reasonably conclude that microfluidic reactor technology offers a series of advantages for the synthesis of solid catalysts. Firstly, the superior micromixing therein provides a (close to) uniform reaction environment for the precipitation process. Secondly, the efficient heat/mass transfer in microreactors renders a (nearly) uniform temperature/concentration distribution [81]. In addition, a fine control over the residence time is attainable in microreactors, which is conducive to the preparation of nanomaterials with a narrow particle size distribution and high repeatability between different flow operation sequences [82,83,84,85]. Moreover, nucleation, growth and agglomeration processes can be effectively controlled by accurately adjusting parameters during reactions in microflow. This unique characteristic enables microreactors to regulate the size, shape, and structure properties of solid catalysts. Finally, it is relatively easy to scale up compared with batch reactors. The scale-up of microreactors is primarily based on the numbering-up concept, that is, the superposition of microchannels. During this scale-up process, the synthesis parameters optimized previously in a single-channel microreactor are kept, so that the transport/reaction characteristics across a multitude of microchannels are identical in principle, which is beneficial to the fast, reliable and modular production of solid catalysts on a large scale [46].

2.3.2. Classification of Microfluidic Methods

According to the flow pattern manipulation during solid catalyst synthesis, microfluidic methods can be divided mainly into two types, namely single-phase flow (continuous flow, Figure 3a) and two-phase flow methods [84]. The two-phase flow method can be subdivided into gas-liquid/liquid-liquid segmented flow (Figure 3b) and liquid-liquid droplet flow (Figure 3c) methods. The length of droplets (if present) in segmented flow is larger than the microchannel diameter and there is usually a film present between the droplet and microchannel wall, while the droplet size is smaller than the microchannel diameter in the droplet flow mode and droplets are surrounded by the continuous liquid phase. In both biphasic flow modes, reactions can occur either in the dispersed phase (Figure 3bI,cI) or in the continuous phase (Figure 3bII,cII). In general, for reactions that are prone to blockage, reagents are usually confined in the dispersed phase, and the generated solid particles are prevented from contacting or interacting with the microchannel, thereby clogging can be prevented. Both single- and two-phase flow methods are characterized by low Reynolds numbers (and thus laminar flow conditions) due to small microchannel sizes, but the mixing behavior of reagents is largely different. The mixing of reagents in microreactors under single-phase flow operation is relatively slow and primarily driven by molecular interdiffusion, whereas reagents under two-phase flow operation can be more efficiently mixed (e.g., on sub-millisecond timescales) through the strong advection generated in droplets [86].

2.3.3. Single-Phase Flow Method

The reagents involved in the single-phase flow method are all miscible (e.g., in commonly used solvents such as water, ethanol and ethylene glycol), and the process of synthesizing solid catalysts is to inject different reagents into microchannels to mix and react. Single-phase flow microfluidic systems for (catalytic) nanomaterial synthesis are considered to be very promising, because this method has the advantages of simple implementation, high flux, and easy control compared with two-phase flow method. However, there are two principal drawbacks that limit the application scope of single-phase flow method. One is that the products deposited on the microreactor wall may cause fouling or even blockage of the microchannel, posing difficulties for long-term operation. The other is that there is a wide distribution of residence time within the microchannel associated with the laminar parabolic velocity profile (Figure 3a), often resulting in a significant dispersion in the particle size and properties of the synthesized solid catalysts.

2.3.4. Two-Phase Flow Method

The two-phase flow method usually uses an inert gas or liquid that does not react or is immiscible with the reagent phase to divide the reagent phase into continuous slugs (i.e., when the wall is wetted by the liquid) or discrete droplets, which can greatly improve the mixing efficiency in the microchannel due to the recirculation of the liquid [87]. This method overcomes the main drawbacks of single-phase flow method, namely, the significant elimination of velocity dispersion and reduction of reactor contamination (due to the confinement of reactions in droplets) [27]. Moreover, the two-phase flow method has the advantages of a large phase boundary area, low axial backmixing degree, and narrow residence time distribution. Therefore, many studies have been focused on the controlled synthesis of nanoparticles using segmented flow and droplet flow methods [88,89].

3. Examples

Microfluidic reactor technology has been extensively used for the synthesis of various catalysts such as metals, metal oxides, and porous materials (zeolite, silica, MOFs). Most of the reports only deal with catalyst synthesis by microfluidic single- or two-phase flow method, and the performance study of catalysts (as listed in Table 2 and Table 3). In addition, some studies adopted different methods to prepare catalysts, and compared the performance of catalysts obtained by different methods and the influence of preparation method on the catalyst properties in detail. In this section, some examples are provided for solid catalysts synthesized by microfluidic methods in recent years. The effects of different synthesis methods on the structural properties and catalytic performance of catalysts are also discussed.

3.1. Metal Catalyst

Metal nanoparticles are widely used in various technological fields, such as catalysis, fine chemicals, petrochemicals, pharmaceuticals, electronics, sensing, fuel cells, and environmental protection due to their excellent physical and chemical properties [90,91]. The size, shape, and structure of metal nanoparticles greatly influence their catalytic, optical, electronic and biomedical properties [92]. Researchers have explored a variety of synthetic strategies for the preparation of metal catalysts with a narrow particle size distribution [93].

3.1.1. Monometallic Catalyst

Among all metals, noble metals (e.g., gold, silver, palladium and platinum) are the most commonly used catalyst materials because of the easy adsorption of reactants on their surfaces, high catalyst activity and selectivity [94]. In the past two decades, the synthesis of gold nanoparticles by microfluidic reactor technology has become a major research hotspot and various gold nanoparticles with controlled sizes and morphologies have been successfully obtained.
As summarized in Table 2, ultrasmall Au nanoparticles with an average diameter of 1.0 nm and narrow size distribution were synthesized in a continuous single-phase flow microfluidic setup using HAuCl4·3H2O as precursor, NaBH4 as reducing agent and PVP as stabilizer [76]. In comparison, the average diameter of Au nanoparticles produced in batch increased to 1.9 nm and the particle size distribution was also slightly broader. Depositing the obtained Au nanoparticles on TiO2 could be used to prepare the catalyst for the oxidation of CO. Au nanoflowers with tunable sizes were synthesized by a rapid continuous single-phase microfluidic approach within 3 min using HAuCl4·4H2O as precursor and HEPES aqueous solution as reducing agent and size regulator (Table 2) [95]. The as-prepared Au nanoflowers exhibited a high catalytic activity for the reduction of 4-nitrophenol. A gas-liquid segmented flow microfluidic reactor was used to synthesize Au nanoparticles (Table 3) with diameters of 3–25 nm at room temperature (using HAuCl4 as precursor and CO as gaseous reductant) [96]. The size and hydrophilicity of Au nanoparticles could be tuned by using different capping agents (Figure 4).
The synthesis of silver nanoparticles has also attracted great research interest because of their special chemical and optical properties [97]. Silver nanoshells with an average shell thickness of 11.6 ± 5.3 nm were successfully synthesized in a single-phase flow central collision-type microreactor using silver-ammonia complexes as precursor, AuNP-SiO2 as core particles and formaldehyde as reducing agent (Table 2) [98]. Compared with the batch synthesis process, the microfluidic synthesis process was time-saving and the obtained shell structures were homogeneous due to the intensive mixing and large surface to volume ratio of the microreactor. Spherical silver nanoparticles with a mean diameter of 7.6 ± 1.8 nm were successfully obtained in a droplet flow microfluidic device using AgNO3 as precursor, NaBH4 as reducing agent, EDTA as complexant and PVP as dispersant (Table 3) [99]. The spherical-shaped Ag nanoparticles with an average size of 4 nm and Au nanoparticles with an average size of 14–17 nm could be obtained in a segmented flow capillary microreactor using AgNO3 and AuHCl4·3H2O as the respective metal precursors, silicon oil as continuous phase, NaBH4 and sodium citrate as reducing agent [100].
Much effort has been taken on the preparation of palladium and platinum nanoparticles with a controllable size and shape due to their wide application in the field of catalysis [101,102]. Pd nanoparticles, which can be used as an electrochemical catalyst, were synthesized in a flow microreactor (under single-phase flow) using PdCl2 as precursor, NaBH4 as reducing agent, and cetyltrimethylammonium bromide as stabilizing agent, and the size of the nanoparticles could be controlled within 5–200 nm by changing the solution flow rate [32]. Ultra-small Pt nanoparticles as well as alloyed nanoparticles including PtPd, PtRu and PtRh (Table 2), which can be used for the catalytic oxidation of volatile organic compounds, were prepared in microreactors (under single-phase flow) using H2PtCl6 (K2PdClO4, RuCl3, RhCl3) as precursor and tetrakis(hydroxymethyl)phosphonium chloride as reducing and stabilizing agent [36]. Compared with the traditional batch process for the synthesis of Pt nanoparticles, the microfluidic method had a short residence time, high yield and high productivity. Pd nanorods with a diameter of about 4 nm and an aspect ratio of up to 6 were prepared by a continuous segmented flow microfluidic method using air as the segmentation gas and Na2PdCl4 mixed in the mixture of water, KBr, ethylene glycol, and polyvinyl pyrrolidone as precursor (Table 3) [103]. This method only required 2 min to synthesize Pd nanorods, while it usually took hours in batch systems. Moreover, the obtained Pd nanorods showed an excellent activity for the catalytic hydrogenation of styrene.

3.1.2. Bimetallic Catalyst

By combining two metals, the properties of the resulting nanoparticles may differ from those of pure metals due to the interaction between metal particles. The bimetallic particles often form alloys or core-shell structures and usually exhibit an excellent catalytic performance because of the synergistic effect of metal atoms [104]. For example, Au@Ag core-shell nanoparticles were synthesized (using HAuCl4 and sodium citrate for the synthesis of Au core, and AgNO3 and ascorbic acid for the preparation of Ag shell) in a single-phase microreactor (Figure 5a) and a conventional batch reactor, respectively [105]. Compared with the batch results, the spherical Au@Ag core-shell nanoparticles synthesized in the microreactor exhibited a more uniform shell size (Figure 5b,c) due to the faster mixing of reagents and better control over the reaction time. Moreover, the microreactor could be used to produce various uniform core-shell nanoparticles, such as Au@Pd (Figure 5d), Au@Au, and Au@Au with an interior nanogap. Hybrid Au-Pd dumbbell-like nanostructures, which involved the synthesis of Pd nanorods using Na2PdCl4 as precursor and ethylene glycol as reducing agent and solvent, as well as the growth of Au on Pd nanorods using HAuCl4 as precursor and ethylene glycol as reducing agent, were successfully synthesized (Table 2) [106]. In more detail, Pd nanorods were synthesized in a sequential-addition microfluidic reactor by segmented flow method, and the growth of Au was conducted in an ultrasonic integrated microfluidic reactor by single-phase flow method. Core@shell Au@Pd nanostructures with controlled shape and dimensions have been obtained in a continuous flow droplet microreactor, including the synthesis of Au core (using HAuCl4·3H2O as precursor, NaBH4 or sodium citrate as reducing agent) and the growth of Pd shell (using H2PdCl4 as precursor, L-ascorbic acid as reducing agent, and silicone oil as the carrier medium) (Table 3) [107]. The droplet-based microreactor further enabled the inline analysis and modification of process conditions compared with the batch method.

3.1.3. Supported Metal Catalyst

Metal nanoparticles used for catalytic reactions usually have a small particle size and high surface free energy, making them easy to agglomerate and difficult to recover and recycle. To overcome these shortcomings, they are usually anchored on solid supports (e.g., carbon materials, metal oxides). The specific surface area of solid supports is usually large, which greatly increases the dispersibility and stability of metal nanoparticles. Moreover, there may be interactions between particles and supports, which will further enhance the catalyst performance [108,109]. Therefore, the design and application of supported metal catalysts is an important method to improve and ensure its catalytic activity and selectivity.
Highly dispersed PtFeCu/C ternary nanocatalysts (Table 2), which showed an excellent catalytic performance for methanol oxidation, were synthesized by single-phase flow method in a microfluidic setup with four parallel channels [110]. The aqueous solution of FeCl3·6H2O, CuCl2·2H2O, and H2PtCl6·6H2O containing suspended carbon black was used as metal precursor solution, and NH3BH3 as reducing agent. The high flow rate and multiple channels increased the throughput obviously. M/TiO2 (M = Pd, Pt or Au) nanoparticles were continuously synthesized by a stabilizer-free method based on a segmented flow microfluidic system using Pd(NO3)2, H2PtCl6 or HAuCl4 mixed with TiO2 aqueous suspension as precursor, NaBH4 with NaOH as reducing agent, and n-octane as continuous phase (Table 3) [77]. Compared with M/TiO2 nanoparticles prepared by the batch method, the obtained nanoparticles in microflow had a smaller noble metal particle size (<5 nm), better dispersity, and thus showed a higher catalytic activity in the hydrolysis of ammonia borane. Ultrafine Ag nanoparticles with a controllable particle size (1.5–5.6 nm) and narrow particle size distribution were uniformly deposited on rGO by segmented flow processing in the microreactor (Figure 6a,c,d), using AgNO3 as precursor, trisodium citrate and NaBH4 as reducing agent, sodium dodecyl sulfate (SDS) as stabilizing agent and octane as continuous phase [111]. The average particle size of Ag nanoparticles in Ag-rGO composites was smaller than those synthesized in batch (Figure 6b).
Table 2. The synthesis parameters, characteristics and application of solid catalysts synthesized by single-phase flow methods in microreactors.
Table 2. The synthesis parameters, characteristics and application of solid catalysts synthesized by single-phase flow methods in microreactors.
Catalytic MaterialMicroreactor
Material a
Reagents bSynthesis ConditionsSize/Morphology cAdvantages of Microfludic MethodTested Catalytic ReactionRef.
AuSilicon/glass HAuCl4·3H2O, NaBH4, PVP13 bar N2 NPs
(1.0 ± 0.4 nm)
Efficient mixing;
narrow size distribution
CO oxidation[76]
AuPTFEHAuCl4·3H2O, trisodium citrate70–100 °C, back pressure at 275 kPaNPs
(1.9 ± 0.2 nm)
Enhanced nucleation rate; inhibited particle growth rate-[112]
AuPDMSHAuCl4·4H2O, HEPES, ammonia 80 μL/min, pH = 7.4NPs
(88 ± 11 nm)
Precise control of reaction timeReduction of 4-nitrophenol[95]
Ag AuNP-SiO2, PVP, AgNO3, HCHO, L-ascorbic acidpH = 8Nanoshells(11.6 ± 5.3 nm)Large surface to volume ratio; rapid mixing [98]
PdPEEK and PTFEH2PdCl4, CTAB, H2Cl6Pt,
L-ascorbic acid
96 °C water bath for 20 min under stirringNanocubes
(≈14 nm)
Efficient mixing and heat transfer;
narrow size distribution
NO2 reduction[113]
PtPTFE tubeH2PtCl6, NaOH, THPC25–95 °Cthermostatic bath, residence time at 1–5 minNPs
(<2 nm)
Short residence time; high yield and productivityn-Hexane oxidation[36]
Au-PdSilicon/glassHAuCl4·3H2O, K2PdCl4, NaBH4, PVP13 bar N2, ice/water bathNPs
(≈1 nm)
Rapid mixingCO oxidation[114]
Au-PdSiliconHAuCl4, KBr, PVP, EG, Na2PdCl4 8 bar, 160 °C, 120 s; 2 bar, 100 °C, 5 minNPs
(15.1 ± 2.3 nm)
Controlled heat/mass transferEthanol oxidation[106]
Pt-Bi Bi(NO3)3·5H2O, H2PtCl6·6H2O, PVP, ethanol, EG, PEGHeat region 200–350 °CNanorods(17 nm)Rapid heat transfer;
precise control of temperature
Methanol fuel
cells
[35]
Pd@AuPtPMMA and PSANa2PdCl4, K2PtCl4, HAuCl4,
L-ascorbic acid, PVP, KBr
60 °C shaking incubator at 200 rpmNPs
(10–12 nm)
High accuracy and reproducibilityH2O2 synthesis[115]
Ag@Cu2OPTFEAgNO3, H2O2, CuSO4·5H2O, SDS,
Na3C6H5O7·2H2O, NaBH4, NaOH
Flow rate at 0.25 mL/min + 0.5 mL/min Ag Cores
(20–50 nm)
High efficiencyMethyl orange degradation[116]
PtFeCu/CStainless steelH2PtCl6·6H2O, FeCl3·6H2O, CuCl2·2H2O, EG, H3N·BH3, Nafion, PEG, carbon black 60 min ultrasonication, 0.5 h stirringNPs
(1.8 ± 0.3 nm)
High throughput Methanol oxidation[110]
Cu-CuOStainless steelCuSO4, NaBH4, PVP, ammonium hydroxide, NaOHpH = 10–12, flow rate at 20 mL/minNanocomposite
(≈10 nm)
Small particle size;
narrow size distribution
Methylene blue degradation[117]
Ag3PO4PDMS and PMMAAgNO3, disodium hydrogen phosphateFlow rate at 130 μL/minNPs
(20–30 nm)
Narrow size distribution;
rapid synthesis
Rhodamine B degradation[118]
TiO2Stainless steelTiOSO4, CO(NH2)2Heating oil 180 °C, back pressure at 2 MPaNPs
(≈5.0 nm)
Small particle size; even size distributionMethylene blue decomposition[119]
a. PTFE: poly(tetrafluoroethylene); PDMS: polydimethylsiloxane; PEEK: polyether ether ketone; PSA: polysulfonamide; PMMA: polymethylmethacrylate. b. PVP: polyvinyl pyrrolidone; HEPES: 2-[4-(2-Hydroxyethyl)-1-piperazinyl]-ethanesulfonic acid; CTAB: cetyltrimethylammonium bromide; THPC: tetrakis(hydroxymethyl)phosphonium chloride; EG: ethylene glycol; PEG: polyethylene glycol; SDS: sodium dodecyl sulfonate. c. NPs: nanoparticles.
Table 3. The synthesis parameters, characteristics and application of solid catalysts synthesized by two-phase flow methods in microreactors.
Table 3. The synthesis parameters, characteristics and application of solid catalysts synthesized by two-phase flow methods in microreactors.
Catalytic MaterialMicroreactor aReagents bSynthesis Conditions cSize/Morphology dAdvantages of Microfluidic MethodTested Catalytic ReactionRef.
Au GlassHAuCl4, CO gas Residence time at 3–5 min, room temperature, atmospheric pressure; SF NPs (3–25 nm)Large surface-to-volume ratio;
high mass transfer
-[96]
Au PTFE and PEEKHAuCl4·3H2O, PVP, PP9
ascorbic acid, C12H16O4
Droplet length at 2.0 mm and distance at 1.33 mm; SFNPs (2.5–4 nm)Fast mixing; low concentration gradient-[120]
Ag PDMS and SiliconAgNO3, NaBH4, N2, PVP, NaOH, NH4OH, EDTA, mineral oilpH = 10–12, room temperature, flow rate at 10–30 mL/h; DFNPs (7.6 ± 1.8 nm)Rapid mixing-[99]
Pd PTFE and silicaNa2PdCl4, KBr, silicone oil, PVP, L-ascorbic acidOil phase flow rate at 180 μL/min, residence time at 6 min; DFNPs (9.0 ± 1.5 nm)Fast and efficient mixing; controlled reaction condition-[121]
Pd Silicon/PyrexNa2PdCl4, PVP, DMF, EG, KBr, air 0.8 MPa, 180 °C, residence time at 10–120 s; SFNanorods (≈4 nm)Reduced synthesis timeStyrene hydrogenation [103]
Ag, AuPTFE and silicaAgNO3, NaBH4, HAuCl4·3H2O, trisodium citrate, silicon oilAg: residence time 0.6 min; Au: oil bath at 105 °C, residence time: 3–20 min; SFNPs (Ag: ≈4 nm;
Au: 16–17 nm)
Controlled residence time-[100]
Au@Pd PTFE and silicaHAuCl4·3H2O, NaBH4, PdCl2, HCl, NaBr, CTAB, CTAC, L-ascorbic acid, sodium citrate, silicone oilTotal flow rate at 0.7 mL/h,
oil bath at 55 °C; DF
Core-shell nanostructures (shell thickness: 2.8 nm)Inline analysis for quick modification of process conditions -[107]
FePtSn/CStainless steel FeCl2·4H2O, NaBH4, H2PtCl6·6H2O,
NMP, PVP
SnCl2·2H2O
Reaction and nucleation at 120 °C, Fe/Pt/Sn ratio at 1/1/0.3, particle growth at 30 °C; DFAlloy NPs (≈2–3 nm; on the carbon surface)Uniform size distributionMethanol oxidation[34]
M/TiO2
(M = Pd, Pt or Au)
PTFE and PFATiO2, NH3BH3, NaBH4,NaOH, Pd(NO3)2·2H2O, H2PtCl6·6H2O, n-octane, HAuCl4·4H2OUltrasonication for 25 min, water bath at 25–80 °C, residence time at 2 min; SFNoble metal NPs (<5 nm; anchored on TiO2)Enhanced mixing; small particle sizeAmmonia borane hydrolysis[77]
Cu, CuOx GlassCuSO4·5H2O, seignette salt, NaOHpH = 12, 350 °C, 2 h; DFMicrospheres (100–500 μm)Controlled particle diameter Hydrogenation; Fenton-like reactions[122]
Ag-rGOPTFE and PFAAgNO3, SDS, octane, NaBH4, trisodium citrate AgNO3:NaBH4 =
1:3 or 1:20; SF
Ag NPs (1.5–5.6 nm)Precise control of reaction parameters [111]
Co/rGO Stainless steel, FEPCoCl2·6H2O, GO, NaBH4, octane Water bath at 40 °C; SF NPs (2.0 ± 0.45 nm)Improved micromixingp-Nitrophenol reduction[123]
Fe3O4 PEEK and PTFEFeSO4·7H2O, KNO3, NaOH, L-lysine, H2SO470–110 °C, 1.4 bar; SFNanoflakes (30 ± 8 nm)Accelerated mixing; reduced processing time-[124]
mSiO2; SiO2@mSiO2; Au@SiO2PTFENH3(aq), ethanol, TEOS, OTMS Gas pressure at 1.5 bar; SFMesoporous
(core-shell) spheres (300–400 nm)
High yield;
controlled particle features
-[125]
Zeolite APTFENaOH, sodium aluminate, sodium silicateOil bath at 90 °C; SF0.9–1.5 µmHigh mixing efficiency;
no clogging
-[126]
MOFPDMS and PFACu(NO3)2·H2O, H3BTC, DMF, ethanol, silicon oilSilicon oil bath at 90 °C; DFOctahedral crystals (5–15 µm)Continuous fabrication; time-saving-[42]
PMMA-polyDADMAC/PtFEPMMA, EGDMA, DADMAC,
AIBN, platinum seed
Water bath at 97 °C; DF≈0.7 µmFast mixing;
short residence time
Azo-dye bleaching[127]
a. PTFE: poly(tetrafluoroethylene); PDMS: polydimethylsiloxane; PEEK: polyether ether ketone; PFA: perfluoroalkoxy alkane; FEP: fluorinated ethylene propylene. b. PVP: polyvinyl pyrrolidone; EG: ethylene glycol; SDS: sodium dodecyl sulfonate; PP9: perfluoromethyldecalin; EDTA: ethylene diamine tetraacetic acid; DMF: dimethyl formamide; CTAB: hexadecyltrimethylammonium bromide; CTAC: cetyltrimethylammonium chloride solution; NMP: N-methyl-2-pyrrolidone; GO: graphene oxide; TEOS: tetraethylorthosilicate; OTMS: trimethoxy silane; MMA: methyl methacrylate; EGDMA: ethylene glycodimethacrylate; DADMAC: diallyldimethyl ammonium chloride; AIBN: azobisisobutyronitrile; H3BTC: 1,3,5- benzenetricarboxylic acid; C12H16O4: 2-hydroxy-4-(2-hydroxyethoxy)-2-methylpropiophenone. c. SF: segmented flow; DF: droplet flow. d. NPs: nanoparticles.

3.2. Metal Oxide Catalyst

Although noble metal catalysts have an excellent catalytic performance, the shortage of resources and high price limit their application scale. Transition metal oxides (e.g., TiO2, ZnO, CeO2, CuO, Fe2O3) are inexpensive and abundant. They have excellent redox properties and can be used as catalysts for many reactions, though their catalytic activity is usually lower than that of noble metal catalysts [15,128]. For example, phosphated TiO2 catalysts were prepared by a continuous sol-gel method in a microreactor (under single-phase flow), which is an eco-friendly catalyst for the selective synthesis of 5-hydroxymethylfurfural (a biobased platform chemical) [12]. Because of the great flexibility and precise control of chemical process, the yield and space-time yield of phosphated TiO2 obtained in the microreactor process were higher than that of the conventional batch synthesis process and the phosphated TiO2 exhibited a remarkably higher glucose conversion and slightly improved selectivity during the catalytic process compared with that obtained in the batch reactor. Benefiting from the high microscopic mixing rate of the membrane dispersion microreactor, the mesoporous γ-Al2O3 nanofibers were synthesized by a simple, template-free and single-phase microfluidic method using NaAlO2 and Al(NO3)3·9H2O as raw materials. And the obtained γ-Al2O3 nanofibers showed considerable pore volumes and uniform pore size distributions compared with those prepared by traditional batch stirring methods [129]. The morphology-controlled synthesis of ZnO nanostructures (spindles, sheets and spheres) was reported via the precipitation of Zn(NO3)2 and NaOH by a facile surfactant-free method in microreactors (under single-phase flow) [130]. These ZnO nanostructures were tested as photocatalysts for the degradation of dye and exhibited an excellent catalytic efficiency.
CeO2 nanoparticles were synthesized through the precipitation of Ce(NO3)3·6H2O and NH3·H2O using a single-phase flow microreactor and a batch stirred tank reactor under the same experimental conditions [39]. CeO2 nanoparticles prepared in the microreactor were nearly 2 times smaller in size, had a better dispersion and narrower particle size distribution than those obtained in batch (Figure 7). When used as catalyst for the mineralization of acid orange 7, the degradation efficiency approached 84.1% for CeO2 nanoparticles synthesized in the microreactor within 0.5 h, which was ~40.6% higher than those prepared in batch. CeO2 nanorods and nanocubes were synthesized by hydrothermal method using Ce(NO3)3·6H2O and NaOH as reagents in a self-made single-phase flow microfluidic system [131]. CeO2 nanorods were transformed to nanocubes by introducing PVP as surfactant under hydrothermal conditions. For comparison, CeO2 nanoparticles were prepared by precipitation method in a stirred tank reactor. The preparation of CeO2 nanostructures by hydrothermal method took only 8 min in the microreactor, while several days were needed in the conventional batch system. The catalytic activity for the decomposition of hydrogen peroxide followed the sequence: CeO2 nanocubes > CeO2 nanorods > CeO2 nanoparticles, which was explained by the microfluidic reaction system that generated nanostructures with a smaller particle size, thereby obtaining larger specific surface areas and better catalytic performance.
In addition, SnO2 nanoparticles were synthesized in a single-phase flow microreactor and a batch reactor using SnCl4·5H2O and NaOH as raw materials to compare its structure and photocatalytic performance [132]. SnO2 nanoparticles obtained in batch were characterized by agglomerated structures composed of extremely tiny nanoparticles with an average size of around 2–3 nm. In contrast, those synthesized in the microreactor had less agglomeration and consisted of extremely small SnO2 nanoparticles with an average size of 3–5 nm. The SnO2 nanoparticles prepared in the microreactor presented photocatalytic activity in the degradation of methylene blue dye because of the significant increase in surface defects. These results demonstrate that the change in synthetic techniques altered the surface properties (e.g., extent of oxygen vacancies and defects at the surface) and optical band gaps, and these changes in turn affected the photocatalytic properties of SnO2 nanoparticles [132].

3.3. Porous Catalytic Material

Porous materials have attracted an extensive scientific research interest in the field of catalysis due to their uniform, abundant, and tunable porous structures, high specific surface area, large adsorption capacity, and excellent physicochemical properties. The most common porous materials that can be used as catalysts include among others zeolites, silica, and MOFs [133].

3.3.1. Silica

Silica is a kind of porous material and its synthesis has received a great deal of research attention due to its wide application in many fields, especially as catalysts and catalyst supports. Over the past decade, microfluidic reactor technology has been introduced to overcome some of the challenges faced in the synthesis of silica, such as precise regulation of its morphology, particle size, and porosity [134]. For instance, silica spheres (diameter: 300–400 nm), core-shell spheres with the mesoporous silica shell (up to ca. 60 nm) and the dense silica core (SiO2@mSiO2) as well as gold-encapsulated spheres (Au@SiO2) were efficiently produced in a microreactor under gas-liquid segmented flow (Table 2; Figure 8a) [125]. Typically, silica spheres were prepared using ammonia as catalyst in the ethanol-water mixture and TEOS in ethanol as precursor. For the silica shell deposition, the mixture of TEOS and OTMS in ethanol was employed as the second precursor. Au@SiO2 particles were synthesized by further adding the aqueous gold colloids in the mixture of ammonia and ethanol. The segmentation of flow was found crucial to obtain the desirable monodispersity of silica spheres or core-shell particles, whereas a significant shell thickness was present under single-phase laminar flow operation.
Hollow silica nanoparticles with an average diameter of ∼30 nm were prepared using tetraethylorthosilicate as precursor, poly (acrylic acid) as template by a micromixer (under single-phase flow) [135]. Compared with the batch reactor, the fast mixing efficiency in the microreactor dramatically reduced the aggregation of poly (acrylic acid) template, which was beneficial to obtain silica nanoparticles with a smaller diameter and narrower size distribution. Silica bearing different surface silanol groups was synthesized using sodium silicate aqueous solution and CO2 as raw materials in a segmented flow microreactor and conventional batch reactor under similar reaction conditions [136]. The obtained silica samples were employed as supports for Pd-based catalysts for benzyl alcohol oxidation. Compared with conventional batch method, silica prepared in the microreactor had a higher specific surface area, more uniform distribution of surface silanol groups and better catalytic effect due to the high mass transfer efficiency of the microreactor.

3.3.2. Zeolite

Zeolite is a kind of crystalline aluminosilicate with regular and uniform pores that has a wide range of applications, especially in the field of catalysis, due to its nanoporosity and acidity [137]. The preparation process of zeolite generally consists of two steps; firstly the mixing of silica and alumina solutions to prepare the hydrogel, and then the hydrothermal crystallization process of the hydrogel. Zeolites with adjustable sizes were prepared in a segmented flow-based microreactor by the one-step hydrothermal synthesis, using silica and alumina solutions as dispersed phase and liquid paraffin as continuous phase (Figure 8b) [126]. The microreactor system had the advantages of no clogging as well as saving time and cost for zeolite synthesis, because of the in situ mixing of silica and alumina solutions (Table 3).
Porous zeolite A nanocrystals with a uniform particle size were efficiently synthesized by a hydrothermal method conducted in a droplet flow-based microreactor, using fluorocarbon oil as continuous phase and the gel solution as dispersed phase (precursor), that is, obtained by the aging of an alumina precursor solution and tetraethylorthosilicate [40]. The aqueous droplets were surrounded by oil medium, preventing the contact of the formed nanocrystals with the microchannel wall, thereby preventing channel clogging. Moreover, it took only tens of minutes to synthesize nanocrystals in the microreactor, compared with hours to days in batch. The obtained zeolite A exhibited a superior catalytic performance and an improved stability in the synthesis of alkyl borate.

3.3.3. MOFs

MOFs, consisting of metal clusters and organic bridging ligands, are a promising class of porous-crystalline materials that can be used in many fields such as catalysis, adsorption, sensing and separation [138]. These materials usually have plenty of pores and high specific surface areas, different transition metals or functional groups, which are very suitable as catalysts. The continuous synthesis of MOFs by means of microreactors has become a research hotspot in recent years due to the ultra-high mixing efficiency and precise control of reaction parameters therein [42].
For example, MIL-88B type MOFs with average sizes of 90–900 nm and narrow particle size distributions were prepared under liquid-liquid segmented flow processing through microreactors, using FeCl3·6H2O solution and three different ligands solution (H2BDC (benzene-1,4-dicarboxylic acid), NH2-H2BDC, Br-H2BDC) as continuous phases and silicone oil as discrete droplets (Figure 8c) [139]. MOFs could be obtained at a residence time as short as 20 s in the microreactor and this method can effectively reduce microreactor clogging by preventing the obtained particles from interacting with the microchannel walls. The synthesis of uniform HKUST-1 microcubes, composed of Cu(II) dimers and 1,3,5-benzenetricarboxylate, was also reported in droplet flow-based microreactors, using silicon oil as continuous phase and the DMSO solution of the precursor of Cu(NO3)2 and 1,3,5-benzenetricarboxylate containing uniform polystyrene nanoparticles as dispersed phase [140].
Figure 8. (a) Schematic of the segmented flow capillary microreactor set-up employed for the synthesis of mSiO2, SiO2@mSiO2, and Au@SiO2 (left) and TEM image of the obtained Au@SiO2 (right). The gas-liquid segmentation present in the microreactor, as well as the progressive SiO2@mSiO2 shell growth, is schematically highlighted in the magnified red rectangle. The inset in the TEM image shows the schematic diagram of the core-shell structure. (b) Schematics of the coaxial capillary assembly and formation of the segments in the PTFE capillary involved in the liquid-liquid segmented flow microfluidic system for one-step continuous synthesis of zeolite A, with the left, middle and right pictures on the bottom illustrating the mixing status of silica and aluminum solutions at the mixing inlet, 1 s and 5 s, respectively (left) and SEM pictures of the obtained zeolites (right). (c) Schematic of the microfluidic setup, segmented flow pattern, and synthesis details of MOFs (left) and the TEM image of Fe-MIL-88B-NH2 synthesized at 95 °C after a residence time of 600 s (right). (ac) were reproduced/adapted with permission from references [125,126,139], respectively.
Figure 8. (a) Schematic of the segmented flow capillary microreactor set-up employed for the synthesis of mSiO2, SiO2@mSiO2, and Au@SiO2 (left) and TEM image of the obtained Au@SiO2 (right). The gas-liquid segmentation present in the microreactor, as well as the progressive SiO2@mSiO2 shell growth, is schematically highlighted in the magnified red rectangle. The inset in the TEM image shows the schematic diagram of the core-shell structure. (b) Schematics of the coaxial capillary assembly and formation of the segments in the PTFE capillary involved in the liquid-liquid segmented flow microfluidic system for one-step continuous synthesis of zeolite A, with the left, middle and right pictures on the bottom illustrating the mixing status of silica and aluminum solutions at the mixing inlet, 1 s and 5 s, respectively (left) and SEM pictures of the obtained zeolites (right). (c) Schematic of the microfluidic setup, segmented flow pattern, and synthesis details of MOFs (left) and the TEM image of Fe-MIL-88B-NH2 synthesized at 95 °C after a residence time of 600 s (right). (ac) were reproduced/adapted with permission from references [125,126,139], respectively.
Powders 01 00011 g008

3.4. Miscellaneous Catalysts

Other kinds of particles (e.g., semiconductors, quantum dots and metal salts) as potential catalyst candidates have also been prepared in microreactors [141]. For example, Ag2S/ZnS core-shell quantum dots (QDs) with a different ZnS shell thickness could be prepared in a droplet microreactor, by interfacial reaction between Zn-oleylamine precursor (in the continuous oil phase) and Ag2S QDs (in the dispersed aqueous phase) [142]. Spherical Ag3PO4 nanoparticles with a narrow size distribution were successfully synthesized in a T-shaped microreactor (via single-phase flow method) using silver nitrate and disodium hydrogen phosphate as raw materials [118]. The photocatalytic efficiency of these Ag3PO4 nanoparticles for the degradation of Rhodamine B dye could reach >97% within only 15 min.

4. Conclusions and Perspectives

The development of microfluidic reactor technology has opened a new and attractive platform for the controlled synthesis of solid catalysts, including among others monometallic and bimetallic catalysts, supported metal catalysts, metal oxides and porous catalytic materials. Compared with solid catalyst synthesis using wet chemical methods (e.g., precipitation, hydrothermal/solvothermal, sol-gel methods as well as thermal decomposition methods) performed in traditional batch reactors, continuous processing in microreactors based on the single-phase or two-phase flow mode could achieve a much better micromixing and heat/mass transfer efficiency. This allowed a (close to) uniform reaction environment, and a fine control over reaction parameters (e.g., concentration, temperature and reaction time), and thus the preparation of solid catalysts with a smaller particle size, more uniform particle size distribution and better morphology control in a significantly shorter reaction time (e.g., typically minutes or even seconds regarding the formation of nanoparticles or nanocrystals) in microreactors than in batch. As a result, solid catalysts prepared by microfluidic approach often exhibited a better catalytic performance in the target applications than their counterparts obtained via batch protocols.
Although a great research progress has been made in the microfludic synthesis of solid catalysts in recent years, there are still considerable scopes for optimization in terms of the synthesis process, catalyst quality, functionality and yield. To advance this technology, some aspects need to be addressed carefully. Firstly, the fouling or even clogging of microchannels may happen by solid products formed in the reaction process because of the agglomeration of particles or the deposition of particles on the microchannel walls due to the gravity or particle-wall interaction. Such phenomena can occur especially during the single-phase flow mode and long-term operation, which might be mitigated via the introduction of strong cavitation effects by coupling ultrasonication with microreactors, the employment of mechanical vibration in continuous flow, the modification of the wettability of microchannel surface (e.g., to be superhydrophobic or superoleophobic), or the use of the two-phase flow method to confine the reaction in the dispersed phase that is not in contact with the wall [143,144]. Secondly, a better synthesis optimization and control can be assisted via coupling microreactors with some online analysis and measurement systems, which enables real-time monitoring and accurate analysis of the as-synthesized particles and immediate adjustment of synthesis conditions to obtain the desired products [19,145]. Thirdly, the majority of current research remains still on the laboratory scale. Effective scale-up strategies for large-scale fabrication of solid catalysts in microreactors with a multitude of microchannels need to be researched further, especially those addressing well the local particle-fluid hydrodynamics [46]. The up-to-date scale-up strategies include the internal numbering-up method (i.e., via arranging a number of identical microchannels in parallel within one microreactor module), the external numbering-up method (i.e., via providing an individual supply system to each microreactor), and the sizing up method (i.e., selective dimension enlarging) [146]. The numbering-up (both internal and external) method may require hundreds to thousands of reaction units to meet high throughput of industrial production. The question of how to achieve a uniform fluid distribution in each microchannel in the microreactor system realized by the internal numbering-up method represents a somewhat significant challenge, especially when it comes to multiphase fluid flow, let alone to the presence of particles [147,148]. In the external numbering-up method, each microreactor requires its own supply system (e.g., pump, buffer tank) and auxiliary equipment (e.g., mixing unit, pressure detector unit, temperature controller unit), resulting in excessive installation and operating costs. The size (e.g., length or diameter) of the microchannel can be increased to some extent to achieve a higher throughput while maintaining the mixing and transfer characteristics at the microscale (the sizing up method). Though this size increase can greatly reduce the occurrence of clogging, the hydrodynamic and transport properties might change dramatically if the channel size becomes too large. Thus, the throughput might be only increased to a somewhat limited extent by this method. To date, a practical and feasible approach for the scale-up of microreactors to satisfy the industrial demand is to combine the strategies of internal numbering-up, external numbering-up, and sizing up, followed by the optimization of the microreactor structures. This approach would facilitate a fast and reliable reproduction of laboratory achievements on the industrial scale [146,147,148,149].
The future research interests are envisaged in the following areas. It is necessary to synthesize more catalytic materials that have not been investigated in microreactors so far, in order to explore the full potential of microfluidic synthesis approach. There exists a lot of room for improvement regarding the activity, selectivity and stability of solid catalysts obtained in microreactors, which necessitates an in-depth fundamental study. Last but not least, the integration of solid catalyst synthesis in microreactors and its further incorporation in microreactors (e.g., in the form of nanofluids, packed beds or catalyst slurries [43,46]) for more sustainable and efficient chemical synthesis represents an emerging area to develop.
To summarize, although microfluidic reactor technology is still in its development phase and faces some challenges, it has become an indispensable tool in the synthesis of solid catalysts. With the continuous and rapid development of this technology, these challenges will be gradually resolved, holding promises for breakthroughs in solid catalyst synthesis. In the near future, microfluidic reactor technology may open new research frontiers and provide unparalleled opportunities for the production of widely used catalysts that not only enable the precise tuning of catalyst particle size, morphology and microstructure, but also catalytic activity, selectivity and other desired functions to fully fulfil the target reaction performance requirements.

Author Contributions

Conceptualization, H.C. and J.Y.; Investigation, H.C.; Writing—original draft preparation, H.C. and Z.D.; Writing—review and editing, J.Y. All authors have read and agreed to the published version of the manuscript.

Funding

The financial support from the Ph.D. Foundation of Henan University of Technology (No. 2018BS047), Natural Science Foundation of Henan Province (No. 202300410108), Colleges and Universities Key Research Program Foundation of Henan Province (No. 20A150013), the Innovative Funds Plan of Henan University of Technology (No. 2020ZKCJ04), and the China Scholarship Council visiting scholarship (No. 202008410370) is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, K.; Li, Y. Catalysis based on nanocrystals with well-defined facets. Angew. Chem. Int. Ed. Engl. 2012, 51, 602–613. [Google Scholar] [CrossRef] [PubMed]
  2. Cuenya, B.R. Synthesis and catalytic properties of metal nanoparticles: Size, shape, support, composition, and oxidation state effects. Thin Solid Film. 2010, 518, 3127–3150. [Google Scholar] [CrossRef]
  3. Ludwig, J.R.; Schindler, C.S. Catalyst: Sustainable catalysis. Chemistry 2017, 2, 313–316. [Google Scholar] [CrossRef]
  4. Hülsey, M.J.; Lim, C.W.; Yan, N. Promoting heterogeneous catalysis beyond catalyst design. Chem. Sci. 2020, 11, 1456–1468. [Google Scholar] [CrossRef] [PubMed]
  5. Chaturvedi, S.; Dave, P.N.; Shah, N. Applications of nano-catalyst in new era. J. Saudi Chem. Soc. 2012, 16, 307–325. [Google Scholar] [CrossRef]
  6. Polshettiwar, V.; Varma, R.S. Green chemistry by nano-catalysis. Green Chem. 2010, 12, 743–754. [Google Scholar] [CrossRef]
  7. Wang, D.; Xie, T.; Li, Y. Nanocrystals: Solution-based synthesis and applications as nanocatalysts. Nano Res. 2009, 2, 30–46. [Google Scholar] [CrossRef]
  8. Deng, X.; Yue, Y.; Gao, Z. Gas-phase photo-oxidation of organic compounds over nanosized TiO2 photocatalysts by various preparations. Appl. Catal. B-Environ. 2002, 39, 135–147. [Google Scholar] [CrossRef]
  9. Bell, A.T. The Impact of Nanoscience on Heterogeneous Catalysis. Science 2003, 299, 1688–1691. [Google Scholar] [CrossRef]
  10. Ghosh Chaudhuri, R.; Paria, S. Core/shell nanoparticles: Classes, properties, synthesis mechanisms, characterization, and applications. Chem. Rev. 2012, 112, 2373–2433. [Google Scholar] [CrossRef]
  11. Solsona, M.; Vollenbroek, J.C.; Tregouet, C.B.M.; Nieuwelink, A.E.; Olthuis, W.; van den Berg, A.; Weckhuysen, B.M.; Odijk, M. Microfluidics and catalyst particles. Lab Chip 2019, 19, 3575–3601. [Google Scholar] [CrossRef]
  12. Martin, O.; Bolzli, N.; Puértolas, B.; Pérez-Ramírez, J.; Riedlberger, P. Preparation of highly active phosphated TiO2 catalysts via continuous sol–gel synthesis in a microreactor. Catal. Sci. Technol. 2019, 9, 4744–4758. [Google Scholar] [CrossRef]
  13. Zhang, F.; Zhang, Y.; Yuan, L.; Gasem, K.A.; Chen, J.; Chiang, F.; Wang, Y.; Fan, M. Synthesis of Cu/Zn/Al/Mg catalysts on methanol production by different precipitation methods. Mol. Catal. 2017, 441, 190–198. [Google Scholar] [CrossRef]
  14. Xue, Q.; Li, Z.; Chen, M.; Wang, Y.; Yan, B.; Luo, G. Co-precipitation continuous synthesis of the Ni-Rh-Ce0.75Zr0.25O2-δ catalyst in the membrane dispersion microreactor system for n-dodecane steam reforming to hydrogen. Fuel 2021, 297, 120785. [Google Scholar] [CrossRef]
  15. Chen, H.; Yang, M.; Tao, S.; Chen, G. Template-free synthesis of Co3O4 nanorings and their catalytic application. CrystEngComm 2018, 20, 679–688. [Google Scholar] [CrossRef]
  16. Chen, H.; Yang, M.; Tao, S.; Ren, M.; Chen, G. Facile synthesis of Co3O4 with different morphologies via oxidation kinetic control and its application in hydrogen peroxide decomposition. Cryst. Growth Des. 2016, 16, 6286–6293. [Google Scholar] [CrossRef]
  17. Shan, W.; Liu, F.; He, H.; Shi, X.; Zhang, C. The Remarkable Improvement of a Ce-Ti based Catalyst for NOx Abatement, Prepared by a Homogeneous Precipitation Method. ChemCatChem 2011, 3, 1286–1289. [Google Scholar] [CrossRef]
  18. Na, H.-S.; Shim, J.-O.; Ahn, S.-Y.; Jang, W.-J.; Jeon, K.-W.; Kim, H.-M.; Lee, Y.-L.; Kim, K.-J.; Roh, H.-S. Effect of precipitation sequence on physicochemical properties of CeO2 support for hydrogen production from low-temperature water-gas shift reaction. Int. J. Hydrogen Energy 2018, 43, 17718–17725. [Google Scholar] [CrossRef]
  19. Phillips, T.W.; Lignos, I.G.; Maceiczyk, R.M.; deMello, A.J.; deMello, J.C. Nanocrystal synthesis in microfluidic reactors: Where next? Lab Chip 2014, 14, 3172–3180. [Google Scholar] [CrossRef]
  20. Shen, J.; Shafiq, M.; Ma, M.; Chen, H. Synthesis and surface engineering of inorganic nanomaterials based on microfluidic technology. Nanomaterials 2020, 10, 1177. [Google Scholar] [CrossRef]
  21. Nakashima, T.; Shimizu, M.; Kukizaki, M. Particle control of emulsion by membrane emulsification and its applications. Adv. Drug. Delivery Rev. 2000, 45, 47–56. [Google Scholar] [CrossRef]
  22. Hakke, V.; Sonawane, S.; Anandan, S.; Sonawane, S.; Ashokkumar, M. Process Intensification Approach Using Microreactors for Synthesizing Nanomaterials—A Critical Review. Nanomaterials 2021, 11, 98. [Google Scholar] [CrossRef] [PubMed]
  23. Yue, J. Green process intensification using microreactor technology for the synthesis of biobased chemicals and fuels. Chem. Eng. Process 2022, 177, 109002. [Google Scholar] [CrossRef]
  24. Whitesides, G.M. The origins and the future of microfluidics. Nature 2006, 442, 368–373. [Google Scholar] [CrossRef] [PubMed]
  25. Niculescu, A.-G.; Chircov, C.; Bîrcă, A.C.; Grumezescu, A.M. Fabrication and applications of microfluidic devices: A review. Int. J. Mol. Sci. 2021, 22, 2011. [Google Scholar] [CrossRef]
  26. Zhao, C.X.; He, L.; Qiao, S.Z.; Middelberg, A. Nanoparticle synthesis in microreactors. Chem. Eng. Sci. 2011, 66, 1463–1479. [Google Scholar] [CrossRef]
  27. Nightingale, A.M.; Demello, J.C. Segmented flow reactors for nanocrystal synthesis. Adv. Mater. 2013, 25, 1813–1821. [Google Scholar] [CrossRef]
  28. Wang, J.; Song, Y. Microfluidic Synthesis of Nanohybrids. Small 2017, 13, 1604084. [Google Scholar] [CrossRef] [PubMed]
  29. Yao, X.; Zhang, Y.; Du, L.; Liu, J.; Yao, J. Review of the applications of microreactors. Renew. Sustain. Energy Rev. 2015, 47, 519–539. [Google Scholar] [CrossRef]
  30. Wagner, J.; Köhler, J. Continuous synthesis of gold nanoparticles in a microreactor. Nano Lett. 2005, 5, 685–691. [Google Scholar] [CrossRef]
  31. Lin, X.Z.; Terepka, A.D.; Yang, H. Synthesis of silver nanoparticles in a continuous flow tubular microreactor. Nano Lett. 2004, 4, 2227–2232. [Google Scholar] [CrossRef]
  32. Sharada, S.; Suryawanshi, P.L.; Kumar, R.; Gumfekar, S.P.; Narsaiah, T.B.; Sonawane, S.H. Synthesis of palladium nanoparticles using continuous flow microreactor. Colloids Surf. A 2016, 498, 297–304. [Google Scholar] [CrossRef]
  33. Hossain, M.J.; Rahman, M.; Rahman, M.; Ali, M.; Nandi, N.; Noor, P.; Ahmed, K.; Akhter, S. Optimized reduction conditions for the microfluidic synthesis of 1.3 ± 0.3 nm Pt clusters. J. Nanostruct. Chem. 2016, 6, 49–56. [Google Scholar] [CrossRef]
  34. Wang, Z.; Fan, H.; Liang, H.; Ma, J.; Li, S.; Song, Y.; Wang, R. Microfluidic synthesis and characterization of FePtSn/C catalysts with enhanced electro-catalytic performance for direct methanol fuel cells. Electrochim. Acta 2017, 230, 245–254. [Google Scholar] [CrossRef]
  35. Zhang, D.; Wu, F.; Peng, M.; Wang, X.; Xia, D.; Guo, G. One-step, facile and ultrafast synthesis of phase-and size-controlled Pt–Bi intermetallic nanocatalysts through continuous-flow microfluidics. J. Am. Chem. Soc. 2015, 137, 6263–6269. [Google Scholar] [CrossRef]
  36. Laura, U.; Arruebo, M.; Sebastian, V. Towards the continuous production of Pt-based heterogeneous catalysts using microfluidic systems. Dalton Trans. 2018, 47, 1693–1702. [Google Scholar] [CrossRef]
  37. Huang, C.; Wang, Y.; Luo, G. Preparation of highly dispersed and small-sized ZnO nanoparticles in a membrane dispersion microreactor and their photocatalytic degradation. Ind. Eng. Chem. Res. 2013, 52, 5683–5690. [Google Scholar] [CrossRef]
  38. Eskandarloo, H.; Zaferani, M.; Kierulf, A.; Abbaspourrad, A. Shape-controlled fabrication of TiO2 hollow shells toward photocatalytic application. Appl. Catal. B-Environ. 2018, 227, 519–529. [Google Scholar] [CrossRef]
  39. Yao, H.; Wang, Y.; Luo, G. A size-controllable precipitation method to prepare CeO2 nanoparticles in a membrane dispersion microreactor. Ind. Eng. Chem. Res. 2017, 56, 4993–4999. [Google Scholar] [CrossRef]
  40. Hoang, P.H.; Yoon, K.-B.; Kim, D.-P. Synthesis of hierarchically porous zeolite A crystals with uniform particle size in a droplet microreactor. RSC Adv. 2012, 2, 5323–5328. [Google Scholar] [CrossRef]
  41. Echaide-Górriz, C.; Clément, C.; Cacho-Bailo, F.; Téllez, C.; Coronas, J. New strategies based on microfluidics for the synthesis of metal–organic frameworks and their membranes. J. Mater. Chem. A 2018, 6, 5485–5506. [Google Scholar] [CrossRef]
  42. Faustini, M.; Kim, J.; Jeong, G.-Y.; Kim, J.Y.; Moon, H.R.; Ahn, W.-S.; Kim, D.-P. Microfluidic approach toward continuous and ultrafast synthesis of metal–organic framework crystals and hetero structures in confined microdroplets. J. Am. Chem. Soc. 2013, 135, 14619–14626. [Google Scholar] [CrossRef] [PubMed]
  43. Yue, J. Multiphase flow processing in microreactors combined with heterogeneous catalysis for efficient and sustainable chemical synthesis. Catal. Today 2018, 308, 3–19. [Google Scholar] [CrossRef]
  44. Frost, C.G.; Mutton, L. Heterogeneous catalytic synthesis using microreactor technology. Green Chem. 2010, 12, 1687–1703. [Google Scholar] [CrossRef]
  45. Hao, N.; Nie, Y.; Zhang, J.X. Microfluidic synthesis of functional inorganic micro-/nanoparticles and applications in biomedical engineering. Int. Mater Rev. 2018, 63, 461–487. [Google Scholar] [CrossRef]
  46. Zong, J.; Yue, J. Continuous Solid Particle Flow in Microreactors for Efficient Chemical Conversion. Ind. Eng. Chem. Res. 2022, 61, 6269–6291. [Google Scholar] [CrossRef]
  47. Makgwane, P.R.; Ray, S.S. Synthesis of nanomaterials by continuous-flow microfluidics: A review. J. Nanosci. Nanotechnol. 2014, 14, 1338–1363. [Google Scholar] [CrossRef] [PubMed]
  48. Leubner, I.H. The Balanced Nucleation and Growth (BNG) Model for Controlled Crystal Nucleation and Size in Controlled Precipitations. In Nucleation and Atmospheric Aerosols; Springer: Dordrecht, The Netherlands, 2007; pp. 48–49. [Google Scholar]
  49. Ring, J. Fundamentals of crystallization: Kinetic effects on particle size distributions and morphology. Chem. Eng. Sci. 1991, 46, 2389–2427. [Google Scholar]
  50. Lindberg, M.; Rasmuson, Å.C. Reaction crystallization in strained fluid films. Chem. Eng. Sci. 2001, 56, 3257–3273. [Google Scholar] [CrossRef]
  51. Ren, M.; Yang, M.; Li, S.; Chen, G.; Yuan, Q. High throughput preparation of magnesium hydroxide flame retardant via microreaction technology. RSC Adv. 2016, 6, 92670–92681. [Google Scholar] [CrossRef]
  52. Koirala, R.; Pratsinis, S.E.; Baiker, A. Synthesis of catalytic materials in flames: Opportunities and challenges. Chem. Soc. Rev. 2016, 45, 3053–3068. [Google Scholar] [CrossRef] [PubMed]
  53. Adam, R.E.; Pozina, G.; Willander, M.; Nur, O. Synthesis of ZnO nanoparticles by co-precipitation method for solar driven photodegradation of Congo red dye at different pH. Photonics Nanostruct. Fundam. Appl. 2018, 32, 11–18. [Google Scholar] [CrossRef]
  54. Feng, Y.; Jiao, T.; Yin, J.; Zhang, L.; Zhang, L.; Zhou, J.; Peng, Q. Facile preparation of carbon nanotube-Cu2O nanocomposites as new catalyst materials for reduction of p-nitrophenol. Nanoscale Res. Lett. 2019, 14, 78. [Google Scholar] [CrossRef] [PubMed]
  55. Shao, J.; Lin, F.; Li, Y.; Tang, H.; Wang, Z.; Liu, P.; Chen, G. Co-precipitation synthesized MnOx-CeO2 mixed oxides for NO oxidation and enhanced resistance to low concentration of SO2 by metal addition. Catalysts 2019, 9, 519. [Google Scholar] [CrossRef]
  56. Paul, B.; Purkayastha, D.D.; Dhar, S.S. One-pot hydrothermal synthesis and characterization of CoFe2O4 nanoparticles and its application as magnetically recoverable catalyst in oxidation of alcohols by periodic acid. Mater. Chem. Phys. 2016, 181, 99–105. [Google Scholar] [CrossRef]
  57. Meeyoo, V.; Panchan, N.; Phongprueksathat, N.; Traitangwong, A.; Guo, X.; Li, C.; Rirksomboon, T. Low temperature methanation of CO2 on high Ni content Ni-Ce-ZrOδ catalysts prepared via one-pot hydrothermal synthesis. Catalysts 2020, 10, 32. [Google Scholar] [CrossRef]
  58. Du, J.; Qu, Z.; Dong, C.; Song, L.; Qin, Y.; Huang, N. Low-temperature abatement of toluene over Mn-Ce oxides catalysts synthesized by a modified hydrothermal approach. Appl. Surf. Sci. 2018, 433, 1025–1035. [Google Scholar] [CrossRef]
  59. Yu, F.; Wang, Y.; Ma, H.; Zhou, M. Hydrothermal synthesis of FeS2 as a highly efficient heterogeneous electro-Fenton catalyst to degrade diclofenac via molecular oxygen effects for Fe (II)/Fe (III) cycle. Sep. Purif. Technol. 2020, 248, 117022. [Google Scholar] [CrossRef]
  60. Yang, B.; Xie, W.; Wang, G.; Wang, Z.; Zhao, X.; Yang, C.; Su, X.; Ma, K. One-step solvothermal synthesis of Co3O4/Co(OH)2 hybrids as a catalyst for visible-light-driven water oxidation. Mater. Lett. 2020, 262, 127170. [Google Scholar] [CrossRef]
  61. Ramakrishnan, V.M.; Natarajan, M.; Santhanam, A.; Asokan, V.; Velauthapillai, D. Size controlled synthesis of TiO2 nanoparticles by modified solvothermal method towards effective photo catalytic and photovoltaic applications. Mater. Res. Bull. 2018, 97, 351–360. [Google Scholar] [CrossRef]
  62. Zhang, X.-F.; Wang, A.-J.; Zhang, L.; Yuan, J.; Li, Z.; Feng, J.-J. Solvothermal synthesis of monodisperse PtCu dodecahedral nanoframes with enhanced catalytic activity and durability for hydrogen evolution reaction. ACS Appl. Energy Mater. 2018, 1, 5054–5061. [Google Scholar] [CrossRef]
  63. Esposito, S. “Traditional” Sol-Gel Chemistry as a Powerful Tool for the Preparation of Supported Metal and Metal Oxide Catalysts. Materials 2019, 12, 668. [Google Scholar] [CrossRef] [PubMed]
  64. Shafei, A.; Sheibani, S. Visible light photocatalytic activity of Cu doped TiO2-CNT nanocomposite powder prepared by sol–gel method. Mater. Res. Bull. 2019, 110, 198–206. [Google Scholar] [CrossRef]
  65. Ren, X.; Yao, Y.; Ren, P.; Wang, Y.; Peng, Y. Facile sol-gel synthesis of C@MoSe2 core-shell composites as advanced hydrogen evolution reaction catalyst. Mater. Lett. 2019, 238, 286–289. [Google Scholar] [CrossRef]
  66. Moghaddam, S.V.; Rezaei, M.; Meshkani, F.; Daroughegi, R. Carbon dioxide methanation over Ni-M/Al2O3 (M: Fe, CO, Zr, La and Cu) catalysts synthesized using the one-pot sol-gel synthesis method. Int. J. Hydrogen Energy 2018, 43, 16522–16533. [Google Scholar] [CrossRef]
  67. Chen, H.; Yang, M.; Tao, S.; Chen, G. Oxygen vacancy enhanced catalytic activity of reduced Co3O4 towards p-nitrophenol reduction. Appl. Catal. B-Environ. 2017, 209, 648–656. [Google Scholar] [CrossRef]
  68. Zheng, J.; Wang, Z.; Chen, Z.; Zuo, S. Mechanism of CeO2 synthesized by thermal decomposition of Ce-MOF and its performance of benzene catalytic combustion. J. Rare Earths 2021, 39, 790–796. [Google Scholar] [CrossRef]
  69. Du, J.; Bao, J.; Liu, Y.; Kim, S.H.; Dionysiou, D.D. Facile preparation of porous Mn/Fe3O4 cubes as peroxymonosulfate activating catalyst for effective bisphenol A degradation. Chem. Eng. J. 2019, 376, 119193. [Google Scholar] [CrossRef]
  70. Asencios, Y.J.; Elias, K.F.; de Zawadzki, A.; Assaf, E.M. Synthesis-Gas Production from Methane over Ni/CeO2 Catalysts Synthesized by Co-Precipitation Method in Different Solvents. Methane 2022, 1, 72–81. [Google Scholar] [CrossRef]
  71. Yu, J.; Yue, L.; Liu, S.; Huang, B.; Zhang, X. Hydrothermal preparation and photocatalytic activity of mesoporous Au–TiO2 nanocomposite microspheres. J. Colloid Interface Sci. 2009, 334, 58–64. [Google Scholar] [CrossRef]
  72. Xu, R.; Wang, D.; Zhang, J.; Li, Y. Shape-dependent catalytic activity of silver nanoparticles for the oxidation of styrene. Chem. Asian J. 2006, 1, 888–893. [Google Scholar] [CrossRef] [PubMed]
  73. Babu, B.; Koutavarapu, R.; Harish, V.; Shim, J.; Yoo, K. Novel in-situ synthesis of Au/SnO2 quantum dots for enhanced visible-light-driven photocatalytic applications. Ceram. Int. 2019, 45, 5743–5750. [Google Scholar] [CrossRef]
  74. Islam, M.R.; Rahman, M.; Farhad, S.; Podder, J. Structural, optical and photocatalysis properties of sol–gel deposited Al-doped ZnO thin films. Surf. Interfaces 2019, 16, 120–126. [Google Scholar] [CrossRef]
  75. Zhao, W.; Yang, S.; Guo, C.; Yang, J.; Liu, Y. One-step fabrication of Fe3O4–Cu nanocomposites: High-efficiency and low-cost catalysts for reduction of 4-nitrophenol. Mater. Chem. Phys. 2021, 260, 124144. [Google Scholar] [CrossRef]
  76. Tofighi, G.; Lichtenberg, H.; Pesek, J.; Sheppard, T.L.; Wang, W.; Schöttner, L.; Rinke, G.; Dittmeyer, R.; Grunwaldt, J.-D. Continuous microfluidic synthesis of colloidal ultrasmall gold nanoparticles: In situ study of the early reaction stages and application for catalysis. React. Chem. Eng. 2017, 2, 876–884. [Google Scholar] [CrossRef]
  77. Luo, L.; Yang, M.; Chen, G. Continuous synthesis of TiO2-supported noble metal nanoparticles and their application in ammonia borane hydrolysis. Chem. Eng. Sci. 2022, 251, 117479. [Google Scholar] [CrossRef]
  78. Abelló, S.; Pérez-Ramírez, J. Tuning Nanomaterials’ Characteristics by a Miniaturized In-Line Dispersion–Precipitation Method: Application to Hydrotalcite Synthesis. Adv. Mater. 2006, 18, 2436–2439. [Google Scholar] [CrossRef]
  79. NaderiNasrabadi, M.; Mortazavi, Y.; Khodadadi, A.A. Highly sensitive and selective Gd2O3-doped SnO2 ethanol sensors synthesized by a high temperature and pressure solvothermal method in a microreactor. Sens. Actuators B 2016, 230, 130–139. [Google Scholar] [CrossRef]
  80. Ling, F.W.; Abdulbari, H.A.; Chin, S.-Y. Synthesis and characteristics of silica nano-particles using modified sol–gel method in microreactor. Mater. Today Proc. 2021, 42, 7. [Google Scholar] [CrossRef]
  81. Marre, S.; Jensen, K.F. Synthesis of micro and nanostructures in microfluidic systems. Chem. Soc. Rev. 2010, 39, 1183–1202. [Google Scholar] [CrossRef]
  82. Song, H.; Chen, D.L.; Ismagilov, R.F. Reactions in droplets in microfluidic channels. Angew. Chem. Int. Ed. Engl. 2006, 45, 7336–7356. [Google Scholar] [CrossRef]
  83. Zhou, P.; He, J.; Huang, L.; Yu, Z.; Zhou, J. Microfluidic High-Throughput Platforms for Discovery of Novel Materials. Nanomaterials 2020, 10, 2514. [Google Scholar] [CrossRef]
  84. Grumezescu, A.M. Nanomaterials Synthesis through Microfluidic Methods: An Updated Overview. Nanomaterials 2021, 11, 864. [Google Scholar]
  85. Li, G.X.; Li, Q.; Cheng, R.; Chen, S. Synthesis of quantum dots based on microfluidic technology. Curr. Opin. Chem. Eng. 2020, 29, 34–41. [Google Scholar] [CrossRef]
  86. Gonidec, M.; Puigmartí-Luis, J. Continuous- versus Segmented-Flow Microfluidic Synthesis in Materials Science. Crystals 2018, 9, 12. [Google Scholar] [CrossRef]
  87. Liu, Z.M.; Yang, Y.; Du, Y.; Pang, Y. Advances in Droplet-Based Microfluidic Technology and Its Applications. Chin. J. Anal. Chem. 2017, 45, 282–296. [Google Scholar] [CrossRef]
  88. He, Y.; Kim, K.J.; Chang, C. Segmented Microfluidic Flow Reactors for Nanomaterial Synthesis. Nanomaterials 2020, 10, 1421. [Google Scholar] [CrossRef]
  89. Liu, L.; Xiang, N.; Ni, Z. Droplet-based microreactor for the production of micro/nano-materials. Electrophoresis 2020, 41, 833–851. [Google Scholar] [CrossRef]
  90. Gao, C.; Lyu, F.; Yin, Y. Encapsulated metal nanoparticles for catalysis. Chem. Rev. 2020, 121, 834–881. [Google Scholar] [CrossRef]
  91. Narayan, N.; Meiyazhagan, A.; Vajtai, R. Metal nanoparticles as green catalysts. Materials 2019, 12, 3602. [Google Scholar] [CrossRef]
  92. Rodrigues, T.S.; da Silva, A.G.; Camargo, P.H. Nanocatalysis by noble metal nanoparticles: Controlled synthesis for the optimization and understanding of activities. J. Mater. Chem. A 2019, 7, 5857–5874. [Google Scholar] [CrossRef]
  93. Jamkhande, P.G.; Ghule, N.W.; Bamer, A.H.; Kalaskar, M.G. Metal nanoparticles synthesis: An overview on methods of preparation, advantages and disadvantages, and applications. J. Drug Deliv. Sci. Technol. 2019, 53, 101174. [Google Scholar] [CrossRef]
  94. Li, X.; Yang, X.; Huang, Y.; Zhang, T.; Liu, B. Supported noble-metal single atoms for heterogeneous catalysis. Adv. Mater. 2019, 31, 1902031. [Google Scholar] [CrossRef]
  95. Fu, Q.; Fu, C.; Teng, L.; Li, W.; Handschuh-Wang, S. Rapid synthesis and growth process deconvolution of Au nanoflowers with ultrahigh catalytic activity based on microfluidics. J. Mater. Sci. 2021, 56, 12. [Google Scholar] [CrossRef]
  96. Huang, H.; du Toit, H.; Ben-Jaber, S.; Wu, G.; Panariello, L.; Thanh, N.T.K.; Parkin, I.P.; Gavriilidis, A. Rapid synthesis of gold nanoparticles with carbon monoxide in a microfluidic segmented flow system. React. Chem. Eng. 2019, 4, 884–890. [Google Scholar] [CrossRef]
  97. Kang, W.J.; Cheng, C.Q.; Li, Z.; Feng, Y.; Shen, G.R.; Du, X.W. Ultrafine Ag nanoparticles as active catalyst for electrocatalytic hydrogen production. ChemCatChem 2019, 11, 5976–5981. [Google Scholar] [CrossRef]
  98. Watanabe, S.; Miyahara, M.T. Flow synthesis of silver nanoshells using a microreactor. Chem. Eng. J. 2019, 374, 674–683. [Google Scholar]
  99. Xu, L.; Peng, J.; Yan, M.; Zhang, D.; Shen, A.Q. Droplet synthesis of silver nanoparticles by a microfluidic device. Chem. Eng. Process. Process Intensif. 2016, 102, 186–193. [Google Scholar] [CrossRef]
  100. Ochoa-Vazquez, G.; Kharisov, B.; Arizmendi-Morquecho, A.; Cario, A.; Aymonier, C.; Marre, S.; López, I. Continuous Segmented-Flow Synthesis of Ag and Au Nanoparticles Using a Low Cost Microfluidic PTFE Tubing Reactor. IEEE Trans. Nanobiosci. 2021, 21, 135–140. [Google Scholar] [CrossRef]
  101. Lin, L.; Yao, S.; Gao, R.; Liang, X.; Yu, Q.; Deng, Y.; Liu, J.; Peng, M.; Jiang, Z.; Li, S. A highly CO-tolerant atomically dispersed Pt catalyst for chemoselective hydrogenation. Nat. Nanotechnol. 2019, 14, 354–361. [Google Scholar] [CrossRef] [PubMed]
  102. Ding, Y.; Wu, Q.; Lin, B.; Guo, Y.; Guo, Y.; Wang, Y.; Wang, L.; Zhan, W. Superior catalytic activity of a Pd catalyst in methane combustion by fine-tuning the phase of ceria-zirconia support. Appl. Catal. B-Environ. 2020, 266, 118631. [Google Scholar] [CrossRef]
  103. Sebastian, V.; Basak, S.; Jensen, K.F. Continuous synthesis of palladium nanorods in oxidative segmented flow. AIChE J. 2016, 62, 373–380. [Google Scholar] [CrossRef]
  104. Huang, J.; Mensi, M.; Oveisi, E.; Mantella, V.; Buonsanti, R. Structural sensitivities in bimetallic catalysts for electrochemical CO2 reduction revealed by Ag–Cu nanodimers. J. Am. Chem. Soc. 2019, 141, 2490–2499. [Google Scholar] [CrossRef]
  105. Shin, Y.; Lim, Y.; Kwak, T.; Hwang, J.H.; Georgescu, A.; Huh, D.; Kim, D.; Kang, T. Microfluidic Multi-Scale Homogeneous Mixing with Uniform Residence Time Distribution for Rapid Production of Various Metal Core–Shell Nanoparticles. Adv. Funct. Mater. 2020, 31, 2007856. [Google Scholar] [CrossRef]
  106. Sebastian, V.; Zaborenko, N.; Gu, L.; Jensen, K.F. Microfluidic assisted synthesis of hybrid Au–pd dumbbell-like nanostructures: Sequential addition of reagents and ultrasonic radiation. Cryst. Growth Des. 2017, 17, 2700–2710. [Google Scholar] [CrossRef]
  107. Santana, J.S.; Koczkur, K.M.; Skrabalak, S.E. Synthesis of Core@ Shell nanostructures in a continuous flow droplet reactor: Controlling structure through relative flow rates. Langmuir 2017, 33, 6054–6061. [Google Scholar] [CrossRef]
  108. Rezaei, F.; Dinari, M. Novel covalent organic polymer-supported Ag nanoparticles as a catalyst for nitroaromatics reduction. Colloids Surf. A 2021, 618, 126441. [Google Scholar] [CrossRef]
  109. Zhang, X.; Chen, K.H.; Zhou, Z.H.; He, L.N. Reduced graphene oxide supported Ag nanoparticles: An efficient catalyst for CO2 conversion at ambient conditions. ChemCatChem 2020, 12, 4825–4830. [Google Scholar] [CrossRef]
  110. Zhou, Y.; Wang, D.; Kang, X.; Zhang, D.; Guo, G. A Scalable Synthesis of Ternary Nanocatalysts for High Efficiency Electrooxidation Catalysis by Microfluidics. Nanoscale 2020, 12, 12647–12654. [Google Scholar] [CrossRef]
  111. Tao, S.; Yang, M.; Chen, H.; Chen, G. Continuous Synthesis of Highly Uniform Noble Metal Nanoparticles over Reduced Graphene Oxide Using Microreactor Technology. ACS Sustain. Chem. Eng. 2018, 6, 8719–8726. [Google Scholar] [CrossRef]
  112. Huang, H.; du Toit, H.; Besenhard, M.O.; Ben-Jaber, S.; Dobson, P.; Parkin, I.; Gavriilidis, A. Continuous flow synthesis of ultrasmall gold nanoparticles in a microreactor using trisodium citrate and their SERS performance. Chem. Eng. Sci. 2018, 189, 422–430. [Google Scholar] [CrossRef]
  113. Pekkari, A.; Say, Z.; Susarrey-Arce, A.; Langhammer, C.; Härelind, H.; Sebastian, V.; Moth-Poulsen, K. Continuous Microfluidic Synthesis of Pd Nanocubes and PdPt Core–Shell Nanoparticles and Their Catalysis of NO2 Reduction. ACS Appl. Mater. Interfaces 2019, 11, 36196–36204. [Google Scholar] [CrossRef] [PubMed]
  114. Tofighi, G.; Gaur, A.; Doronkin, D.E.; Lichtenberg, H.; Wang, W.; Wang, D.; Rinke, G.N.; Ewinger, A.; Dittmeyer, R.; Grunwaldt, J.-D. Microfluidic synthesis of ultrasmall AuPd nanoparticles with a homogeneously mixed alloy structure in fast continuous flow for catalytic applications. J. Phys. Chem. C 2018, 122, 1721–1731. [Google Scholar] [CrossRef]
  115. Van Nguyen, H.; Kim, K.Y.; Nam, H.; Lee, S.Y.; Yu, T.; Seo, T.S. Centrifugal microfluidic device for the high-throughput synthesis of Pd@ AuPt core–shell nanoparticles to evaluate the performance of hydrogen peroxide generation. Lab Chip 2020, 20, 3293–3301. [Google Scholar] [CrossRef]
  116. Tao, S.; Yang, M.; Chen, H.; Ren, M.; Chen, G. Microfluidic synthesis of Ag@Cu2O core-shell nanoparticles with enhanced photocatalytic activity. J. Colloid. Interface Sci. 2017, 486, 16–26. [Google Scholar] [CrossRef] [PubMed]
  117. Xu, L.; Srinivasakannan, C.; Peng, J.; Zhang, L.; Zhang, D. Synthesis of Cu-CuO nanocomposite in microreactor and its application to photocatalytic degradation. J. Alloys Compd. 2017, 695, 263–269. [Google Scholar] [CrossRef]
  118. Singh, A.; Baruah, A.; Katoch, V.; Vaghasiya, K.; Prakash, B.; Ganguli, A.K. Continuous flow synthesis of Ag3PO4 nanoparticles with greater photostability and photocatalytic dye degradation efficiency. J. Photochem. Photobiol. A 2018, 364, 382–389. [Google Scholar] [CrossRef]
  119. Deng, N.; Wang, Y.; Luo, G. A novel method for fast and continuous preparation of superfine titanium dioxide nanoparticles in microfluidic system. Particuology 2022, 60, 61–67. [Google Scholar] [CrossRef]
  120. Hafermann, L.; Michael Köhler, J. Small gold nanoparticles formed by rapid photochemical flow-through synthesis using microfluid segment technique. J. Nanopart. Res. 2015, 17, 99. [Google Scholar] [CrossRef]
  121. Kim, Y.H.; Zhang, L.; Yu, T.; Jin, M.; Qin, D.; Xia, Y. Droplet-based microreactors for continuous production of palladium nanocrystals with controlled sizes and shapes. Small 2013, 9, 3462–3467. [Google Scholar] [CrossRef]
  122. Ding, Y.; Yan, Y.; Wang, H.; Wang, X.; Hu, T.; Tao, S.; Li, G. Preparation of Hollow Cu and CuOx Microspheres with a Hierarchical Structure for Heterogeneous Catalysis. ACS Appl. Mater. Interfaces 2018, 10, 41793–41801. [Google Scholar] [CrossRef]
  123. Yang, M.; Luo, L.; Chen, G. Microfluidic synthesis of ultrasmall Co nanoparticles over reduced graphene oxide and their catalytic properties. AIChE J. 2020, 66, e16950. [Google Scholar] [CrossRef]
  124. Larrea, A.; Sebastian, V.; Ibarra, A.; Arruebo, M.; Santamaria, J. Gas slug microfluidics: A unique tool for ultrafast, highly controlled growth of iron oxide nanostructures. Chem. Mater. 2015, 27, 4254–4260. [Google Scholar] [CrossRef]
  125. Knossalla, J.; Mezzavilla, S.; Schüth, F. Continuous synthesis of nanostructured silica based materials in a gas–liquid segmented flow tubular reactor. New J. Chem. 2016, 40, 4361–4366. [Google Scholar] [CrossRef]
  126. Yu, L.; Pan, Y.; Wang, C.; Zhang, L. A two-phase segmented microfluidic technique for one-step continuous versatile preparation of zeolites. Chem. Eng. J. 2013, 219, 78–85. [Google Scholar] [CrossRef]
  127. Li, X.; Visaveliya, N.; Hafermann, L.; Gross, G.A.; Knauer, A.; Koehler, J.M. Hierarchically structured particles for micro flow catalysis. Chem. Eng. J. 2017, 326, 1058–1065. [Google Scholar] [CrossRef]
  128. Nayak, R.; Ali, F.A.; Mishra, D.K.; Ray, D.; Aswal, V.; Sahoo, S.K.; Nanda, B. Fabrication of CuO nanoparticle: An efficient catalyst utilized for sensing and degradation of phenol. J. Mater. Res. Technol. 2020, 9, 11045–11059. [Google Scholar] [CrossRef]
  129. Li, F.; Wan, L.; Wang, Y.; Wang, Y. Template-free method for the synthesis of high-pore-volume γ-Al2O3 nanofibers in a membrane dispersion microreactor. Nanotechnology 2021, 32, 185601. [Google Scholar] [CrossRef]
  130. Baruah, A.; Jindal, A.; Acharya, C.; Prakash, B.; Basu, S.; Ganguli, A.K. Microfluidic reactors for the morphology controlled synthesis and photocatalytic study of ZnO nanostructures. J. Micromech. Microeng. 2017, 27, 035013. [Google Scholar] [CrossRef]
  131. Yao, H.; Wang, Y.; Jing, Y.; Luo, G. Ultrafast, Continuous and Shape-Controlled Preparation of CeO2 Nanostructures: Nanorods and Nanocubes in a Microfluidic System. Ind. Eng. Chem. Res. 2018, 57, 7525–7532. [Google Scholar] [CrossRef]
  132. Periyasamy, M.; Saha, A.; Sain, S.; Mandal, M.; Kar, A. A comparative structural and photocatalytic study on SnO2 nanoparticles fabricated in batch reactor and microreactor. J. Environ. Chem. Eng. 2020, 8, 104604. [Google Scholar] [CrossRef]
  133. Liang, J.; Liang, Z.; Zou, R.; Zhao, Y. Heterogeneous catalysis in zeolites, mesoporous silica, and metal–organic frameworks. Adv. Mater. 2017, 29, 1701139. [Google Scholar] [CrossRef] [PubMed]
  134. Hao, N.; Nie, Y.; Zhang, J. Microfluidics for Silica Biomaterials Synthesis: Opportunities and Challenges. Biomater. Sci. 2019, 7, 2218–2240. [Google Scholar] [CrossRef]
  135. He, Y.; Kim, K.-J.; Chang, C.-H. Continuous, size and shape-control synthesis of hollow silica nanoparticles enabled by a microreactor-assisted rapid mixing process. Nanotechnology 2017, 28, 235602. [Google Scholar] [CrossRef] [PubMed]
  136. Wang, J.; Guo, Z.; Zhao, J.; Yang, Q.; Dai, Y.; Yang, Y.; Wang, C. Preparation of silica as catalyst supports with controlled surface property using continuous flow reactor. Appl. Catal. A 2019, 585, 117212. [Google Scholar] [CrossRef]
  137. Wang, Y.; Wang, C.; Wang, L.; Wang, L.; Xiao, F.-S. Zeolite fixed metal nanoparticles: New perspective in catalysis. Acc. Chem. Res. 2021, 54, 2579–2590. [Google Scholar] [CrossRef]
  138. Yang, J.; Yang, Y.W. Metal–organic frameworks for biomedical applications. Small 2020, 16, 1906846. [Google Scholar] [CrossRef]
  139. Paseta, L.; Seoane, B.; Julve, D.; Sebastián, V.; Téllez, C.; Coronas, J. Accelerating the controlled synthesis of metal–organic frameworks by a microfluidic approach: A nanoliter continuous reactor. ACS Appl. Mater. Interfaces 2013, 5, 9405–9410. [Google Scholar] [CrossRef]
  140. Cui, J.; Gao, N.; Yin, X.; Zhang, W.; Liang, Y.; Tian, L.; Zhou, K.; Wang, S.; Li, G. Microfluidic synthesis of uniform single-crystalline MOF microcubes with a hierarchical porous structure. Nanoscale 2018, 10, 9192–9198. [Google Scholar] [CrossRef]
  141. Campbell, Z.S.; Bateni, F.; Volk, A.A.; Abdel-Latif, K.; Abolhasani, M. Microfluidic synthesis of semiconductor materials: Toward accelerated materials development in flow. Part. Part. Syst. Charact. 2020, 37, 2000256. [Google Scholar] [CrossRef]
  142. Zeng, Y.M.; Pan, L.J.; Wang, J.; Fan, Y.L.; Shu, Y.; Pang, D.W.; Zhang, Z.L. Interfacial synthesis of Ag2S/ZnS core/shell quantum dots in a droplet microreactor. ChemistrySelect 2020, 5, 5889–5894. [Google Scholar] [CrossRef]
  143. Delacour, C.; Lutz, C.; Kuhn, S. Pulsed ultrasound for temperature control and clogging prevention in micro-reactors. Ultrason. Sonochem. 2019, 55, 67–74. [Google Scholar] [CrossRef] [PubMed]
  144. Ryan, L.H. Managing solids in microreactors for the upstream continuous processing of fine chemicals. Org. Process Res. Dev. 2012, 16, 870–887. [Google Scholar]
  145. Suryawanshi, P.L.; Gumfekar, S.P.; Bhanvase, B.A.; Sonawane, S.H.; Pimplapure, M.S. A review on microreactors: Reactor fabrication, design, and cutting-edge applications. Chem. Eng. Sci. 2018, 189, 431–448. [Google Scholar] [CrossRef]
  146. Dong, Z.Y.; Wen, Z.H.; Zhao, F.; Simon, K.; Timothy, N. Scale-up of micro- and milli-reactors: An overview of strategies, design principles and applications. Chem. Eng. Sci. 2021, 10, 100097. [Google Scholar] [CrossRef]
  147. Peng, Z.; Wang, G.; Moghtaderi, B.; Doroodchi, E. A review of microreactors based on slurry Taylor (segmented) flow. Chem. Eng. Sci. 2022, 247, 117040. [Google Scholar] [CrossRef]
  148. Zhang, J.; Wang, K.; Andrew, R.T.; Klavs, F.J.; Luo, G.S. Design and scaling up of microchemical systems: A review. Annu. Rev. Chem. Biomol. Eng. 2017, 8, 285–305. [Google Scholar] [CrossRef] [PubMed]
  149. Wang, K.; Lu, Y.C.; Luo, G.S. Strategy for scaling-up of a microsieve dispersion reactor. Chem. Eng. Technol. 2014, 37, 2116–2122. [Google Scholar] [CrossRef]
Figure 1. The reaction process between M and N ions in the solution to form the solid catalyst MN. The red balls are for M ions, orange balls are for N ions, and small blue balls are for the MN insoluble. The reunited blue balls are for the MN crystal nuclei, and the finally obtained large blue balls are for the target product MN.
Figure 1. The reaction process between M and N ions in the solution to form the solid catalyst MN. The red balls are for M ions, orange balls are for N ions, and small blue balls are for the MN insoluble. The reunited blue balls are for the MN crystal nuclei, and the finally obtained large blue balls are for the target product MN.
Powders 01 00011 g001
Figure 3. Schematic illustration of microfluidic methods for solid catalyst synthesis: (a) single-phase flow (continuous flow) method, characterized by the presence of a laminar parabolic flow velocity; (b,c) two-phase flow method, which can be subdivided into segmented flow method (b) and droplet flow method (c). Adapted with permission from reference [19].
Figure 3. Schematic illustration of microfluidic methods for solid catalyst synthesis: (a) single-phase flow (continuous flow) method, characterized by the presence of a laminar parabolic flow velocity; (b,c) two-phase flow method, which can be subdivided into segmented flow method (b) and droplet flow method (c). Adapted with permission from reference [19].
Powders 01 00011 g003
Figure 4. TEM images of gold nanoparticles capped by (a) trisodium citrate, (b) TWEEN 80, (c) oleylamine in octane and (d) thiol-PEG-COOH. Inserts in the upper right corner of (a,b) show the flow patterns in the microreactor arranged in a configuration of coiled flow inverter. Inserts in the bottom of (ad) show the values of average particle size and polydispersity of gold nanoparticles. Reproduced with permission from reference [96].
Figure 4. TEM images of gold nanoparticles capped by (a) trisodium citrate, (b) TWEEN 80, (c) oleylamine in octane and (d) thiol-PEG-COOH. Inserts in the upper right corner of (a,b) show the flow patterns in the microreactor arranged in a configuration of coiled flow inverter. Inserts in the bottom of (ad) show the values of average particle size and polydispersity of gold nanoparticles. Reproduced with permission from reference [96].
Powders 01 00011 g004
Figure 5. (a) Schematic diagram of the microfluidic device for the facile production of various uniform metal core-shell nanoparticles. The microreactor with built-in upper staggered herringbone microstructures had two inlets (fed with two aqueous solutions) and one outlet. The right image shows the upper herringbone structure of the microchannel from the top view and side view, and the mixing and flow of nanoparticles. TEM images of Au@Ag core-shell nanoparticles: (b) microfluidic synthesis; (c) batch synthesis. TEM images of Au@Pd core-shell nanoparticles: (d) microfluidic synthesis; (e) batch synthesis. Reproduced with permission from reference [105].
Figure 5. (a) Schematic diagram of the microfluidic device for the facile production of various uniform metal core-shell nanoparticles. The microreactor with built-in upper staggered herringbone microstructures had two inlets (fed with two aqueous solutions) and one outlet. The right image shows the upper herringbone structure of the microchannel from the top view and side view, and the mixing and flow of nanoparticles. TEM images of Au@Ag core-shell nanoparticles: (b) microfluidic synthesis; (c) batch synthesis. TEM images of Au@Pd core-shell nanoparticles: (d) microfluidic synthesis; (e) batch synthesis. Reproduced with permission from reference [105].
Powders 01 00011 g005
Figure 6. (a) The microreactor system for the synthesis of Ag-rGO composites. The system can be divided into two stages: (1) the deposition of Ag nanoparticles onto GO sheets, (2) the reduction of Ag-GO composites to Ag-rGO composites; (b) TEM images of Ag-rGO synthesized in the batch reactor; (c,d) TEM images of Ag-rGO synthesized in the microreactor at relatively low (c) and high (d) Ag loadings. Reproduced with permission from reference [111].
Figure 6. (a) The microreactor system for the synthesis of Ag-rGO composites. The system can be divided into two stages: (1) the deposition of Ag nanoparticles onto GO sheets, (2) the reduction of Ag-GO composites to Ag-rGO composites; (b) TEM images of Ag-rGO synthesized in the batch reactor; (c,d) TEM images of Ag-rGO synthesized in the microreactor at relatively low (c) and high (d) Ag loadings. Reproduced with permission from reference [111].
Powders 01 00011 g006
Figure 7. The membrane dispersion-based microreactor system for CeO2 nanoparticle preparation, with the shown flow profile and inner structure of the microreactor (a); TEM images and primary particle size distribution of CeO2 particles synthesized in the microreactor (b,c) and batch stirred tank reactor (d,e). Reproduced/adapted with permission from reference [39].
Figure 7. The membrane dispersion-based microreactor system for CeO2 nanoparticle preparation, with the shown flow profile and inner structure of the microreactor (a); TEM images and primary particle size distribution of CeO2 particles synthesized in the microreactor (b,c) and batch stirred tank reactor (d,e). Reproduced/adapted with permission from reference [39].
Powders 01 00011 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, H.; Dong, Z.; Yue, J. Advances in Microfluidic Synthesis of Solid Catalysts. Powders 2022, 1, 155-183. https://doi.org/10.3390/powders1030011

AMA Style

Chen H, Dong Z, Yue J. Advances in Microfluidic Synthesis of Solid Catalysts. Powders. 2022; 1(3):155-183. https://doi.org/10.3390/powders1030011

Chicago/Turabian Style

Chen, Huihui, Zhenhua Dong, and Jun Yue. 2022. "Advances in Microfluidic Synthesis of Solid Catalysts" Powders 1, no. 3: 155-183. https://doi.org/10.3390/powders1030011

Article Metrics

Back to TopTop